Next Article in Journal
Performance of a Nanofluid-Cooled Segmented Thermoelectric Generator: Hollow/Filled Leg Structures and Segmentation Effects
Previous Article in Journal
The Use of Response Surface Methodology to Optimize Assisted Extraction of Bioactive Compounds from Cucurbita maxima Fruit By-Products
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Impacting Parameters for Photocatalytic Degradation of Organic Effluents by Ferrites and Their Nanocomposites

1
Department of Chemistry, Punjab Agricultural University, Ludhiana 141001, Punjab, India
2
Department of Soil Science, Punjab Agricultural University, Ludhiana 141001, Punjab, India
3
Department of Soil and Water Engineering, Punjab Agricultural University, Ludhiana 141001, Punjab, India
*
Author to whom correspondence should be addressed.
Processes 2023, 11(6), 1727; https://doi.org/10.3390/pr11061727
Submission received: 30 April 2023 / Revised: 21 May 2023 / Accepted: 28 May 2023 / Published: 5 June 2023
(This article belongs to the Section Catalysis Enhanced Processes)

Abstract

:
Traditional wastewater treatment methods, such as reverse osmosis, adsorption, desalination, and others, are outweighed by the photocatalytic degradation of organic pollutants. Ferrites are prominent photocatalysts due to their tunable band gaps, surface areas, and magnetic properties, which render photodegradation economical. Ferrites and their nanocomposites have been reported as promising visible light active photocatalysts. The photocatalytic system is heavily reliant on a number of factors that influence the photodegradation of organic effluents. This review demonstrates various parameters such as substrate concentration, pH of solution, photocatalyst quantity, photocatalyst surface area, metal and non-metal ion doping, light intensity, irradiation time, quenchers, etc. affecting the photocatalytic degradation of organic effluents by ferrite nanoparticles and their nanocomposites in detail. The photodegradation efficiency of the ferrite nanoparticles alters with the change in the value of pH of the solution, which further depends upon the nature of the pollutant used. A dose of the substrate and the photocatalyst must be optimized so as to attain better photodegradation efficiency. Photocatalysts with different surface areas change the amount of active sites, which in turn affects the degradation of pollutant and render it a crucial factor. In addition, the mechanism of the action of photocatalysis is elaborated in this review. Future research perspectives for the advancement of ferrites and their nanocomposites are deliberated in order to improve their use as photocatalysts.

Graphical Abstract

1. Introduction

Organic pollutants in water sources have increased across the world due to the continuous growth of global industry and population. Nearby ecosystems are being poisoned by the toxic and carcinogenic effluents from the pharmaceutical, agricultural, and textile industries, for example [1,2]. These effluents greatly affect groundwater quality, aquatic life, the environment, and human health [3,4,5,6,7]. There are numerous techniques for water treatment, including electrodialysis, electrochemical methods, adsorption, coagulation, bioremediation, and ultrafiltration [8,9,10,11,12,13,14,15,16,17]. However, these methods have limitations to the effective purification of wastewaters and investments [18,19].
Advanced oxidation processes (AOPs) are quite effective for wastewater remediation. Reducing the concentration of harmful compounds in wastewater is the primary goal of AOPs before its discharge. To generate high-potential chemical oxidants on-site, ozone (O3), UV/visible light, hydrogen peroxide (H2O2), and Fenton’s reagent (Fe(II)/H2O2) can be used [20]. Photocatalysis is one of the renowned AOPs because photocatalysts can utilize UV/visible or solar energy to degrade pollutants, making photocatalysis economically feasible [21,22]. Photocatalysis is a process whereby a catalyst initiates or stimulates a chemical reaction using light. This mechanism is referred to as a “photocatalytic reaction” [23,24]. For treating wastewater-containing organic pollutants and dyes, heterogeneous photocatalysis using AOPs is an excellent option because it uses renewable solar energy. Therefore, it is strongly recommended to eliminate organic pollutants and colours [25]. Various reactive species, such as hydroxyl radicals (HO.), superoxide anion (O2), holes (h+), and singlet oxygen (O21), degrade contaminants adsorbed on the surface of a photocatalyst [26]. Apart from these oxidants, other solution parameters such as the pH of the solution [27], photocatalyst quantity [28], photocatalyst surface area [29], metal and non-metal ion doping [30], light exposure, irradiation time [31], and quenchers [28] also influence the photodegradation of organic pollutants.
Ferrites and their composites—because of their effective adsorption efficiency, excellent visible light catalytic activity, and outstanding Fenton reaction potential—have been regarded as promising materials for the removal of contaminants such as heavy metal/metalloids and organics [32]. Due to the decrease in particle size due to nanoscale confinement and surface effects, the magnetic properties of ferrite nanoparticles display peculiar characteristics that differ from those of bulk equivalent materials [33]. The advantages of magnetic ferrite nanoparticles include their affordability and excellent physical and chemical stability [34]. The magnetic, electric, and optical properties of the matter are affected by the behaviour of the ferrites in the nanoscale range, which is dominated by the quantum effect [35].
As of April 2023, search results on the Science Direct database for the phrase “photocatalytic activity of ferrite” in the titles of publications, keywords, and abstracts of the reviews are displayed in Figure 1. The publications on pure, doped, and composite (binary, ternary, and quaternary) MFe2O4 are included in the search results. There was a sizable rise in the number of publications between 2015 and 2023. Due to their magnetic properties, simplicity in synthesis, availability of precursors, and biocompatibility, ferrites have received extensive research interest in the field of the photocatalytic degradation of pollutants. The impact of various parameters on the photodegradation of organic pollutants using ferrite and their nanocomposites as photocatalysts, however, is not discussed in any of the reviews. Focusing on these issues, this review provides a thorough discussion of each parameter with regard to the photodegradation of the reported organic pollutants, a brief overview of the synthesis of ferrites and their nanocomposites, and a discussion of the photodegradation mechanism of organic pollutants. Finally, conclusions are drawn and the future research areas that still require attention are listed.

2. Structure of Ferrites

Ferrites are considered prominent photocatalysts because of their tunable band gaps, stability, distinctive magnetic properties, and low cost of production. In general, the chemical formula of ferrites is AB2O4. The tetrahedral (A) and octahedral (B) sites are occupied by metal cations, and the arrangement of the oxygen anions is cubic closed packing, as shown in Figure 2. This structure can provide different combinations of metal cations due to the balanced negative charges of oxygen anions. When all the A sites are occupied by divalent metal cations and the B sites are occupied by iron (III) cations, the structure that is obtained is called a normal spinel structure, having the formula MeA[Fe2]BO4. Inverse spinel ferrites are structures in which an equal number of iron (III) ions occupy both the sites and all the B sites are occupied by divalent metal ions and have the formula FeA[MeFe]BO4. Some metal cations strongly prefer a particular site, such as Zn for the A site.Thus, ZnFe2O4 is a normal spinel [36,37,38].
The perfect structure of the mixed ferrite crystals mentioned above, however, are affected by variations in the synthesis conditions, and instead have an empirical formula of [M1−xFex]A[MxFe2−x]BO4 (0≤ x≤ 1). The addition of a third metal creates ternary [38], with complexity in the arrangement of the metal ions. Ferrites are found to have different electrical, optical, and magnetic properties, as well as different activities in some catalytic reactions, depending on the composition and arrangement of the metal ions [36].

3. Methods of Synthesis of Ferrites and Their Nanocomposites

There are many different chemical and physical processes used for the synthesis of ferrite nanoparticles. Chemical-based techniques have been used to synthesise Fe2O3 nanoparticles, including reverse micelles [40], co-precipitation [41], sonochemical reactions [42], thermal decomposition [43], sol-gel synthesis [44], hydrothermal [45], and micro-emulsion technology [46].
Ball milling is a mechanical method for solid-state synthesis and is frequently used to grind powders into extremely tiny particles. Being a cost-effective and eco-friendly approach, it is mostly used in industries for large-scale production. However, the size and shape of ferrite nanoparticles are non-uniform. Regulation of the preparative parameters is required, including concentration, pH, type of complexing agent, temperature, etc. Ferrites are typically produced at a temperature of several hundred degrees. At higher temperatures, it is challenging to deposit film ferrites of a consistent quality. In addition, due to the release of poisonous gases during deposition, high-temperature processes are more susceptible to environmental pollution. Oxidation of metallic substrates and corrosion are prevented using low temperature deposition techniques. Table 1 represents the advantages and disadvantages of various physical and chemical methods used in the synthesis of ferrites and their nanocomposites.
Among various chemical methods used in the synthesis of ferrites, different methods have their own advantages and disadvantages. For example, the solvothermal method can be used to control the particle size of the synthesized material [47,48]. The sol-gel method is one of the best methods as it yields ferrite with more surface area of the nanoparticles, which is useful for the degradation of pollutants [49,50,51,52].
However, due to their magnetic nature and strong magnetic dipole interactions, magnetic NPs are particularly susceptible to agglomeration [53]. The agglomeration observed in these NPs also restricts their use in various applications [53,54,55,56,57,58,59]. To overcome these problems, different techniques have been proposed, such as doping and the fabrication of composites using wet chemical methods (hydrothermal, solvothermal), sol-gel, co-precipitation, combustion, and ultrasonication methods. Various ferrites which have been synthesized using these methods and their use as photocatalysts in the degradation of various pollutants is discussed in Table 2.
Hollow microspheres of Ce3+-doped NiCo-ferrites were synthesized using a carbon sphere as a template in a template-based deposition and surface reaction method [60]. The hydrothermal method showed the particle size of carbon sphere as 0.5 µm, while the template-based method showed the particle size of Ni0.5Co0.5Fe2O4 as 300 nm. Graphene-based CuCe0.2Fel.8O4 nanocomposites were synthesized using the hydrothermal method [61,62]. Furthermore, on examining the morphology, thermal stability, and crystal structure of CuFe2O4, CuCe0.2Fel.8O4 ferrite NPs, and graphene-based CuCe0.2Fel.8O4, the graphene was found to have distinct effects on the crystallite size, surface morphology, lattice strain, and thermal stability of the CuCe0.2Fel.8O4 ferrite NPs.
Nitrogen-doped 3D magnetic graphene gels were synthesized through one-pot hydrothermal synthesis, and it was found that the reduction in GO, N doping process, and generation of Fe3O4 particles occur simultaneously and are accompanied by 3D self-assembly [63]. Nitrogen-doped magnetic graphene gels were considered organized frameworks of graphene sheets with a uniform dispersion of the Fe3O4 NPs. The wet-chemical method is a synthetic route that deals with the chemical reactions in the solution phase using precursors at proper experimental conditions. The spinel ferrites were prepared through a chemical reaction between aqueous solutions of metal nitrates, ferric nitrates, and oxalyl dihydrazide [64].
Table 2. Photocatalytic activity of various ferrite NPs.
Table 2. Photocatalytic activity of various ferrite NPs.
S.No.CompoundMethodSize
(nm)
Band Gap
(ev)
Surface AreaPollutantCatalyst Amount (mg)Degradation %geRef
1.ZnFe2O4Solvothermal
Microwave
Sol-gel
77
40
34.8
1.65
1.9
1.6
51.81
6.0
-
RhB
MB
OI
80
60
400
100
32
36.8
[65]
[66]
[67,68]
2.MnFe2O4Microwave391.56-IC5094[69]
3.CaFe2O4Solution combustion1001.8279.3MB50100[70]
4.MgFe2O4Solid state-25MB5090[71]
5.CuFe2O4Co-precipitation
Solution combustion
26

8
1.92

1.9
-

-
4-CP

MG
10

40
62

82
[72]

[73]
6.NiFe2O4Solution combustion
Co-precipitation
14

29
2

1.78
-

13.48
MG

MO
40

100
94.5

83.2
[73]

[74]
7.BaFe2O4Sol-gel---BY
BV
MB
AO
10,000100[75]
8.CoFe2O4Precipitation
Solvothermal
Sol-gel
17
40
40
3.05
2.33
1.15
-
-
-
MO
MB
RhB
100
50
50
77
93
78
[76]
[77]
[78]
9.CdFe2O4Pyrolysis-2.0615.27CS3055.5[79]
10.SrFe2O4Sol-gel-1.75-RhB2089.3[80]
RhB—Rhodamine B, MB—Methylene blue, OI—Orange I, IC—Indigo Carmine, 4-CP—4- Chlorophenol, MG—Malachite green, MO—Methyl orange, BY—Basic Yellow, BV—Basic Violet, AO—Acid Orange, CS—Ceftiofur sodium.
In the ultrasonication method, the sound energy is applied to agitate the particles in the solution, which causes physical and chemical transformations for the preparation of nanocomposites [81]. Various studies have demonstrated that this approach is the simplest way to synthesize ferrite-based nanocomposites due to its low-cost and use of distilled water rather than other solvents. The nanocomposite of the carbon quantum dots with trimetallic strontium–titanium ferrite synthesized via an ultrasonication approach possesses a high-specific surface area (101.3 m2g−1) anda low band gap coupled with fluorescence-quenching, which facilitated the transfer of the photo-induced charge carriers. This feature is very useful for the photodegradation of nitroaromatics [82]. Hierarchical nanoflowers of MgFe2O4, bentonite and boron, phosphorous co-doped graphene oxide synthesized using the ultrasonication method showed that chiffon-like co-doped graphene oxide nanosheets were wrapped on the MgFe2O4-bentonite surface, resulting in a porous flower-like morphology. The excellent conductivity of the co-doped nanosheets allowed for a quick charge carrier transport, which aided in the separation of those carriers during the photocatalytic reaction [83]. Cationic surfactant cetyltrimethyl ammonium bromide was chosen to coat the MgFe2O4 because its negatively charged surface can interact with the surfactant during ultrasonication [84]. Similarly, the nanocomposite of calcium ferrite with nitrogen-doped graphene oxide (CaFe2O4-NGO) with magnetic properties was synthesized using this approach has been effectively used for the degradation of p-nitrophenol [85].
Apart from the ultrasonication method, the sol-gel approach was also used in the synthesis of ferrite-based nanocomposites, and additional material was added during the synthesis of ferrite. MgFe2O4-bentonite nanocomposites were synthesized in different w/w ratios using this method for the degradation of the p-nitrophenol [86]. The optical band gap of the MgFe2O4 NPs is 2.58 eV, and this decreases to 2.26 eV with increasing bentonite content in nanocomposites due to the presence of Fermi levels. A similar nanocomposite was synthesized for the degradation of martius yellow. It was found that this nanocomposite showed better degradation for malathion as compared to the p-nitrophenol and martius yellow [87]. This may be due to the resonance stabilization of the structure of the martius yellow [86,88].
Cobalt ferrite/reduced GO nanocomposites were prepared using the membrane-solvothermal method [89]. Scanning electron micrographs of nanocomposite showed the presence of most of the rod-shaped CoFe2O4 scattered beside the graphene, and only a few CoFe2O4 nanorods successfully loaded on the graphene. GO-supported ferrite (MFe2O4, M = Fe, Co and Ni) nanocomposites were prepared using the solvothermal method [90]. The uniform size (20–30 nm) of MFe2O4 microspheres was seen in the transmission electron micrograph. Vibrational sample magnetometry studies revealed a superparamagnetic behavior of all the nanocomposites with Ms values between 41 and 43 emug−1 at room temperature. The removal of the dye pollution was found to be useful with the help of the nanocomposites.
The co-precipitation method is widely used as a liquid phase method in the synthesis of ferrite-based nanocomposites. The co-precipitation and hydrothermal methods were used in the synthesis of core-shell ZnFe2O4/ZnS composites [91]. A characterization of the synthesized composite showed the presence of a layer of ZnS clusters on the surface of the ZnFe2O4 NPs. The band gap energy was observed to be 2.2 eV, showing the possibility of its application as a photocatalyst under visible light irradiation. The relative energy band positions in the ZnFe2O4 and ZnS were considered a main advantage for the improved separation of electron-hole pairs. Moreover, the easy magnetic collection of the composite materials was due to the sufficient magnetization of the ZnFe2O4 core. Thus, the different methods induce different properties in the ferrite-based nanocomposites, and each property has its own advantage for the photocatalytic degradation of organic pollutants.

4. Impacting Parameters for Photocatalytic Activity of Ferrites and Their Nanocomposites

The photocatalytic activity of ferrite NPs and their nanocomposites is reliant on various parameters, such as the substrate concentration, pH of the solution, photocatalyst quantity, surface area of the photocatalyst, metal and non-metal ion doping, light intensity, irradiation time, quenchers, etc. The impact of each parameter on the photodegradation of organic effluents is discussed in detail in the subsequent subsections.

4.1. Effect of pH

Solution pH is an important variable in wastewater treatment as it influences the photocatalytic degradation of organic compounds. The pH of heterogeneous catalysts changes the surface charge, and, as a result, the photocatalytic activity of the catalyst changes [92,93].
The surface charge of the catalyst plays a vital role for the photodegradation of anionic dyes, which was determined using spinel NixZn1−xFe2O4 (x = 0.0 to 1.0) NPs [29]. The maximum degradation was achieved at pH 3, due to the electrostatic desirability among the anionic dye and the positively charged surface. At basic pH values, the surface of the catalyst became negatively charged, which caused an electrostatic repulsion between the anionic dye and catalyst, thus retarding the degradation of the dyes (Figure 3a). Similar observations were recorded during the photocatalysis of the methylene blue dye using bismuth ferrite (BiFeO3) NPs, where the dye was completely degraded under visible light irradiation for 80 min [94]. Further, it was reported that sunlight and the presence of a photocatalyst are essential for the dye degradation process. Dye degradation was found to be 53% under dark conditions with the presence of only a catalyst. The dye removal efficiency was negligible in the absence of a catalyst and light. After irradiation under visible light with neutral and acidic pH conditions, no decrease in the concentration of dye was observed. This means that the excitation of dye molecules by photon and then the dispersion of their excitation energy had no role in the decolorization of the dye. However, at a basic pH, the rate of photolysis was high due to the generation of hydroxyl radicals through the generation of hydroxyl ions which degraded the methylene blue dye. This study was supported using an electron spin resonance spectroscopy [95]. The coating of the titania–silica (TiO2–SiO2) on the cobalt ferrite (CoFe2O4) NPs changes the properties on the photocatalyst surface. The SiO2 layer between the CoFe2O4 and TiO2 may prevent the charge injection from the semiconductor TiO2 to the magnetic CoFe2O4, which in turn led to the increase in photocatalytic activity towards the methylene blue dye. Under the acidic conditions, the charge on the TiO2 surface was positive, which provided electrostatic interactions for anionic dye molecules [96,97]. The removal of the congo red using Ni0.6Co0.4Fe2O4 was found to be at a maximum for pH 3 (40.28%) and decreased gradually up to pH 10 (15.45%), which was supported by the isoelectric point value of the dye [98].
The Ag/AgBr/NiFe2O4 plasmonic photocatalyst exhibited a superior photocatalytic performance in the photodecomposition of the cationic Rhodamine B dye under visible light irradiation. The zero-point charge (pHzpc) of the catalyst was about 5.2. Thus, the composite was negatively charged above that pH and provided electrostatic attractions for the cationic dye molecules [99]. The Mn0.5Zn0.125Mg0.375Fe2O4 NPs were stabilized at a basic pH, having negative zeta potential (−75.14 eV), which was very important for the photocatalytic degradation of the cationic Rhodamine B dye, as shown in Figure 3b [100]. The ciprofloxacin and remazol deep red dye were degraded using MRuxFe2−xO4 (M = Ni, Cu) [101]. The dye degradation efficiency was observed to be 93% after 60 min at pH 2.5, which decreased at higher pH levels. In addition, H2O2 was used to degrade the ciprofloxacin without changing the pH of the pure solution. The ciprofloxacin degraded to its greatest extent over a 120-min period at pH 2.5.
Acetamiprid and imidacloprid both belong to the neonicotenoid class of pesticides, but showed different optimum pH values for their maximum degradation using Fe3O4/GO-Fe3O4/CoFe2O4/GO-CoFe2O4and CoFe2O4photocatalyst, respectively. In case of acetamiprid degradation, maximum degradation was achieved at pH 3 [102]. At basic pH values, a decline in photocatalytic activity was observed because of ferrous excess hydroxyl ions reactions. This reaction led to the formation of yellowish-brown ferric hydroxide sludge, which ultimately covered the active sites of the catalyst and hence contributed to the reduction in overall efficiency. The degradation potentials of the catalyst (in all cases) dropped when the pH approached basic values. In the case of the imidacloprid, maximum degradation was achieved at pH > 7.0, which may be attributed to the formation of the hydroxide radicals on the surface of ZnO/CoFe2O4 magnetic nanocomposite, leading to the degradation of the imidacloprid pesticide [103].
In case of the phenolic compounds, such as phenol, 2-nitrophenol, 2-chlorophenol, p-chlorophenol, 2, 4, and 6-tricholorophenol, maximum degradation occurred in the pH range of 5 to 11, using ammonia modified graphene-Fe3O4, ZnFe2O4 multi-porous microbricks/graphene, Fe3O4@β-CD, and Fe3O4-gC3N4 nanocomposites [104,105,106,107]. Under alkaline conditions, the interaction between the phenolic compounds and photocatalyst is retarded due to the instability of the oxygen functionalities of the graphene sheets, which are deprotonated and dissolved in the aqueous medium. Contrary results were obtained for the maximum degradation of the phenol (94.2%), which was achieved at pH 3.5, using manganese ferrite [108]. This may be due to the presence of peroxydisulfate, as its adsorption on the positively charged catalyst surface in the pH range of 3–6 was more intense due to its negatively charged SO4¯·radical. In the pH range of 3–7, efficiencies greater than 90% were attained, but at values of pH up to 9, efficiency decreased to 65%.
Thus, the pH of the solution plays a crucial role in the degradation of pollutants. However, the optimum pH for the removal of the organic effluents using ferrites and their nanocomposites as photocatalysts depends upon the type of photocatalyst and the type of organic effluent used.
Figure 3. (a) Effect of pH on the degradation of methylene blue dye using NixZn1−xFe2O4 (x = 0.0 to 1.0) NPs. Adapted with permission [29]; copyright (2016), Elsevier. (b) Effect of pH on the degradation of Rhodamine B dye using Mn0.5Zn0.125Mg0.375Fe2O4 NPs. Adapted with permission [100]; copyright (2020), Elsevier.
Figure 3. (a) Effect of pH on the degradation of methylene blue dye using NixZn1−xFe2O4 (x = 0.0 to 1.0) NPs. Adapted with permission [29]; copyright (2016), Elsevier. (b) Effect of pH on the degradation of Rhodamine B dye using Mn0.5Zn0.125Mg0.375Fe2O4 NPs. Adapted with permission [100]; copyright (2020), Elsevier.
Processes 11 01727 g003

4.2. Effect of Catalyst Dose

The dose of ferrites used in the photodegradation process is another important parameter. Various studies have reported that the photocatalyst dose is highly related to the degradation of the organic pollutants during photocatalysis [84,109].
The methylene blue dye degradation increased upto a certain level when the magnetic catalyst Ni0.6Zn0.4Fe2O4 loading was increased [29,110]. Furthermore, the negative influence of the catalyst loading on the degradation was observed. The initial rise in the degradation was due to the increase in the number of the active sites present on the surface of the catalyst, increasing the production of ·OH and O2−· radicals. Similar results were reported for the degradation of the methylene blue using MFe2O4 (M = Co, Ni, Cu, Zn) and ZnFe2O4, respectively [111,112]. The decrease after a certain time was due to either the aggregation of catalyst particles at high concentrations, causing a decrease in the number of surface active sites, or an increase in the opacity and light scattering of the catalyst particles at high concentrations, leading to a decrease. The maximum photodegradation of the methyl orange using ZnSm1.5Fe0.5O4 photocatalyst was achieved with 0.1 g/100 mL of catalyst. Unfavourable light scattering and a decrease in the light flow decreased the efficiency of the degradation, even on increasing the amount of catalysts [113]. The ciprofloxacin and remazol deep red dye were degraded using Ru-doped nanoferrites and as a function of the catalyst dosage. The percentage degradation was found to be at a maximum at 0.50 gL−1 ferrite concentration in both cases. The availability of more catalytic sites was expected to increase the degradation efficiency with increasing catalyst amounts; however, a large amount of ferrite causes turbidity of the solution, which prevents light from reaching the reaction medium and tends to impair the degradation process [101]. Using the Mn0.75Co0.25Fe2O NPs, the effectiveness of the catalyst concentration towards methylene blue degradation after UV light exposure was investigated. The catalyst dose was varied from 10 to 30 mg. It was found that the Ct/C0 value decreased from 0.14 to 0.05, indicating an improvement in the degradation performance [114]. On increasing the Mn0.5Zn0.125Mg0.375Fe2O4 photocatalyst dose, it was found that the value of the Ct/C0 declines from 60 to 50, corresponding to 40% to 50% Rhodamine B removal after 80 min. The higher doses of the catalyst led to an increased absorption of photons by the catalyst, and the degradation efficiency improved (Figure 4a).
Similar behavior has been observed while using CoFe2O4/rGO composite as a photocatalyst for the degradation of methylene blue dye [115]. When the catalyst dosage was increased upto 0.5 gL−1, the degradation efficiency increased to 47% due to the increased absorption of photons on the surface of the photocatalyst (Figure 4b). However, overdoses of the sample (1 gL−1) led to falls in the degradation effectiveness. Barium ferrite for the degradation of basic yellow 28, methylene blue, acid orange 7, and basic violet showed similar increases with increases in the catalyst loading from 1 to 10 gL−1 [76]. The influence of the catalyst dose on the degradation efficiency of acetamiprid was studied by varying the catalyst dose from 0 to 200 mg L−1. On increasing the amount of GO-Fe3O4and GO-CoFe2O4 upto certain limit, the degradation increases because of the availability of more catalyst sites for the absorption of light [102]. Effective degradation was achieved for each of the four catalysts, at a catalyst dose of 100 mg L−1, and the trend of degrading the acetamiprid was observed increase. However, a further increase in the catalyst concentration led to decreased degradation rates because of the agglomeration of the catalyst. CuCe-Ferrite/TiO2 was used for the photodegradation of the reactive red 250 dye, and the efficiency was found to be 100% using 0.32 g/100 mL of the catalyst. A shift to more wavelength was observed due to the combination of CuCe-Ferrite andTiO2 and hence the absorption of more visible light [28].
Thus, the photocatalyst dose is an important factor that affects the efficiency of a photocatalyst. In general, it is observed that increasing the photocatalyst dose can lead to a higher efficiency of photo degradation. However, excessively high concentrations can lead to the aggregation of ferrites, which can reduce their efficiency [84].

4.3. Effect of Surface Area of Photocatalyst

The photocatalytic activity of the nanomaterials is enhanced when their surface area is large. The surface area of the nanoparticles is increased with a decrease in particle size, ultimately improving their dispersion in a solution and enhancing their interaction with pollutants. This leads to an improvement in their photocatalytic activity.
The effect of the surface areas of the magnetic NPs system NixZn1−xFe2O4, with different concentrations of Ni2+ ions (0.0 ≤ x ≤ 1.0), on the photocatalytic degradation of methylene blue was studied [29]. The Ni0.6Zn0.4Fe2O4was found to have better photocatalytic activity, having a surface area of (36.6 m2g−1) [110]. The particle size and morphology of the Cu1−xMnxFe2O4 effects the photocatalytic degradation of the 4-chlorophenol [116]. For the photodegradation of pendimethalin, core-shell nanocomposites of Sr0.4Ti0.6Fe2O4.6@SiO2 were used. The increased specific surface area and more active sites observed in the scanning electron and transmission electron micrographs were attributed to the higher photocatalytic efficiency in the core-shell nanocomposites. The highest percentage photodegradation of the pendimethalin was observed in the Sr0.4Ti0.6Fe2O4.6@SiO2 core-shell nanocomposites, which was found to be 96% [117]. The study on the influence of Zn content on the efficiency of Zn/Fe2O4 degradation in visible light radiation [118] reported that there was a 40% residual concentration of carbamazepine after an 80-min iron oxide irradiation. On the other hand, the residual carbamazepine concentration decreased as the zinc content increased (Figure 5). Since photocatalytic degradation is a surface-mediated process, the improved degradation efficiency is the result of the increased specific surface area of the zinc-rich nanoparticles [119].
The calcination temperature also plays vital role in the increase of the surface area of the photocatalyst. It was observed that on decreasing the calcination temperature of MgFe2O4 from 600 °C to 500 °C, the particle size of the NPs increased which in turn increased the surface area and ultimately enhanced the photocatalytic activity [120]. The removal of the basic yellow 28, methylene blue, acid orange 7, and basic violet has been studied using barium ferrite composite [76]. The complete decomposition of the dye was observed using 1000 °C sintered catalyst, while catalysts that were sintered at 700 °C or 800 °C degraded only 30% of the dye. The superior removal efficiency of the catalyst sintered at 1000 °C was due to presence of more barium ferrite phase. On the contrary, the degradation of the congo red, methyl orange, methylene blue, Rhodamine B, 4-nitrophenol, and 4-chlorophenol using the CoFe2O4 xerogel was found to decrease from 96% to 75% when it was calcined at a temperature range of 500–900 °C. As the calcining temperature rose, the rate of degradation of the nano-CoFe2O4 photocatalyst decreased, which was related to the BET-specific surface area of the CoFe2O4 NPs [121].
Hence, the particle size, surface area, and availability of the active sites are very important characteristics for enhancing photocatalytic properties. Small particle sizes with high crystallinity are crucial in this regard because their higher specific surface area encourages greater photocatalytic activity. Sometimes, synthesizing at a high temperature is necessary to create particles with a high crystallinity, but it also increases the particle size. As a result, one must consider how to maximize crystallinity without sacrificing particle size [122].

4.4. Effect of Substrate Concentration

The dependence of the substrate concentration on the photocatalytic reaction rate is an important factor. The photocatalytic degradation was the highest for 10 mg·L−1 of dye concentration, which decreased on further increasing the methyl orange concentration from 10 to 30 mg L−1 due to absorption of solar light by the dye [113]. A contradictory behavior was observed while using the CoFe2O4/rGO photocatalyst for the methylene blue [115]. The photocatalytic efficiency of the CoFe2O4/rGO photocatalyst increased from 47 to 55% on raising the initial dye concentration from 2 to 5 mg·L−1 and decreased to 45% for 15 mg·L−1. The strong interactions between the nanocomposite and dye molecules were responsible for the initial rise in the degradation efficiency (Figure 6). However, lesser photons adsorption by the photocatalyst due to the increase in the methylene blue dye decreased the activity of the catalyst. Similar behavior was observed in MFe2O4 (M = Co, Ni, Cu, Zn) for the degradation of the methylene blue upto 94% using 20 mg·L−1 of photocatalyst. The efficiency decreased to 65% for a 50 mg·L–1 methylene blue solution. However, the limited production of the active radicals, which were insufficient to degrade the higher concentration of the methylene blue dye, was the main factor in the outcome. The photogeneration of the electron-hole pairs was also hampered by the highly concentrated methylene blue dye solution, as it reduced photon penetration in the solution phase [111]. Similar results were observed during the degradation of the methylene blue dye using CuFe2O4/TiO2 and ZnFe2O4 [109,112].
The effect of the substrate concentration on the photodegradation of reactive red 250 dye using TiO2 and CuCe-Ferrite/TiO2 NPs has been examined [28]. When the concentration of the reactive red 250 was increased to 100 ppm, a decrease in the chemical oxygen demand was observed over TiO2 and CuCe-Ferrite/TiO2 NPs. These observations were correlated to a decrease in the light penetration because an increase in the reactive red 250 dye concentration resulted in the incorporation of additional reactive red 250 dye molecules onto the TiO2 and CuCe-Ferrite/TiO2 NPs. The Fe3O4, GOFe3O4, CoFe2O4, and GO-CoFe2O4 effectively (80%) degraded 10 mg·L−1 solution of acetamiprid. The degradation efficiency was observed to be 100% on increasing the acetamiprid from 2 to 8 mg·L−1. However, as the pesticide dose was increased from 10 mg·L−1 to 16 mg·L−1, more active sites were occupied due to the increased addition of pollutant molecules [102].

4.5. Effect of Irradiation Intensity and Time

The intensity of the irradiation is another important factor responsible for the photodegradation process. The effect of the irradiation intensity has been studied for the photocatalytic degradation of basic yellow 28, methylene blue, acid orange 7, basic violet using barium ferrite, and rGO composite [76]. With an increase in the irradiation intensity from 3500 to 15,000 lx, the degradation efficiency went from 3 to 25%. The degradation rose to 47.4% with an additional increase in the irradiation intensity to 26,000 lx. The clear indication that the catalyst activity was based on photocatalysis rather than any other assumed mechanisms was due to the sizeable increase in degradation with irradiation intensity. At higher irradiation intensities, the catalyst surface was exposed to more photons, which can be used to fuel the degradation process.
The effect of irradiation time has been studied for the photodegradation of methylene blue using CoFe2O4/rGO. The degradation efficiency increased on increasing the irradiation time to 180 min with catalyst dosage of 0.5 gL−1, and a methylene blue dye concentration of 5 mg·L−1 reached a relatively stable value [115] (Figure 4b and Figure 6). ZnFe2O4 NPs were used to measure the rate of the methylene blue dye decomposition over time. When the dye was exposed to UV light, the degradation increased (99.8%) with longer exposure times, up to 45 min. A plateau was reached after 45 min because the degradation did not increase [112].
A rapid increase in the degradation of acetamiprid was observed when the irradiation time was varied from 15 to 120 min and studied at an ideal pH, oxidant dose, dose of pesticide, and catalyst. The rate slowed down with time, and the active sites required 60 min to completely degrade the pollutant [102]. When the magnesium ions were doped in the cobalt ferrite NPs at various fractions, the rate of photo degradation increased and the degradation time fell to 225 min, which was 330 min for the pure cobalt ferrite. An increase in the Mg2+ concentration decreases the rate of degradation due to the consequent increase in the number of active sites on the surface of the photocatalyst [123].

4.6. Effect of Doping

The doping of ferrites is also emerging as a new method for enhancing photocatalytic potential, as doped ferrites display enhanced visible light absorption and improved photodegradation efficiency towards various types of organic pollutants [122].
The effectiveness of MRuxFe2−xO4 (M = Ni, Cu) NPs was assessed for remazol deep red and ciprofloxacin photocatalytic degradation [101]. With an increasing Ru content, the catalytic activity of the MRuxFe2−xO4 (x = 0, 0.02, 0.06, 0.1, and 0.4) increased. The NiRu0.4Fe1.6O4 demonstrated the best catalytic performance, with complete degradation in 5 min. When the Ru was substituted, the increased catalytic activity supported the preference for the octahedral site. The CuRuxFe2−xO4 showed an increase in the photocatalytic efficiency, which was found to be 92% for x = 0.02, whereas with further substitution up to x = 0.06 and x = 0.1, the degradation time reached 120 min and 90 min, respectively. Similarly, when the Ti content in the ternary-doped Sr1−xTixFe2O4+δ ferrite was increased from 0.3 to 0.6, the band gap was decreased from 2.45 to 2.18 eV, which increased the charge carrier concentration and decreased the barrier height from 1.20 to 1.02 eV, compared to the pure SrFe2O4 and TiFe2O5 [83]. These features were very helpful for the degradation of the p-nitrophenol, pendimethalin and martius yellow with apparent rate constant values 1.9 × 10−2 min−1, 2.3 × 10−2 min−1, and 1.3 × 10−2 min−1, respectively. Similar results have been reported for the degradation of rhodamine B dye [124].
Furthermore, doping Y3+ into pure CoFe2O4has improved the stability of the composite and increased photocatalytic activity compared to that of pure CoFe2O4 NPs [125]. The doping increased the stability and effectiveness of the CoFe2O4 NPs when exposed to visible light. However, higher concentrations of metal ions may cause the photocatalyst to lose its crystallite and have more internal defects. As a result, it decreased the efficiency of the photocatalytic system. Moreover, the photocatalytic activity was increased by using rare earth metals such as Dy3+, Nd3+, Eu3+, Gd3+, and others as dopants [126]. The photocatalytic efficiency increased, while the energy band gap decreased (1.95 to 1.42) when the Sm3+ was doped into the pure ZnFe2O4 (ZnSmxFe2−XO4). High photocatalytic degradation of methyl orange was observed when the ZnSm1.5Fe0.5O4 NPs were used as a photocatalyst. After 80 min, 92.3% of the degradation was observed [113].
Generally, the doping of ferrites causes a narrowing of the band gap, which aids in in the absorption of lower energy photons and trapping of electrons [127,128,129]. This renders them a superior visible light active photocatalyst compared to the other nanomaterials. However, it is difficult to predict the best kind of dopant, as their efficiency can change with changes in other operational parameters.

4.7. Effect of Quenchers

In order to conclude the probable photocatalytic degradation mechanism of different organic pollutants over the photocatalyst surface, various quenchers have been added to study the quenching effect. It has been observed that the use of various scavengers decreases the photocatalytic degradation efficiencies [28]. Different quenchers, such as ethylene diamine tetraacetate (EDTA, h+ scavenger), methanol/butanol (CH3OH/C4H9OH, OH scavenger), ascorbic acid (C6H8O6, O2− scavenger), and sodium azide (NaN3, O2* species) work according to the following reactions:
Processes 11 01727 i001
Ascorbic acid acts as an electron donor and converts superoxide radical (O2−) into hydrogen peroxide (H2O2).
Processes 11 01727 i002
Hydroxyl radical (HO·) abstracts hydrogen from methanol and converted into formaldehyde. In this way, methanol quenches hydroxyl radicals.
Processes 11 01727 i003
Singlet oxygen forms charge transfer complex with N3 which ultimately decomposes into N3 radical.
Various studies based on photocatalysis have demonstrated that out of ethylene diamine tetraacetate, methanol/butanol, ascorbic acid, and sodium azide, ascorbic acid possesses the highest inhibiting effect, which means that the O2−radicals were primary species responsible for the degradation of the organic pollutants [84,117,130].The addition of ascorbic acid greatly quenched (61.8%) the photodegradation of the martius yellow, followed by the butanol (73.7%), ethylene diamine tetraacetate (92.3%) and methanol (90.9%), using magnesium ferrite-bentonite as a photocatalyst [88]. This might be due to the increase in the number of the hydroxyl radicals in the presence of hydrogen peroxide, a potent oxidising agent [114]. A similar quenching effect was studied for methyl orange and tetracycline degradation [131].
However, a limited number of papers have been published addressing the effect of quenchers on the degradation of organic pollutants. Thus, clear insight about the influence of quenchers has not been explored yet.

4.8. Photodegradation Mechanism

Photocatalysis using ferrite-based nanomaterials as photocatalysts involves the interaction of photons of light to breakdown various organic pollutants [132,133,134,135,136]. They consist of semiconductors with large bandgaps where electrons are readily promoted from valence into the conduction band after the absorption of photons. This results in the generation of photoexcited electrons (e) and positive-valence band holes (h+). The photocatalytic mechanism of degradation works through pollutants being reduced or oxidized by photoexcited electrons (e) and holes (h+). In water, hydroxide (OH) ions (generated when positive-valence band holes (h+) are trapped by water) are non-selective strong oxidizing agents that oxidize organic pollutants to water and gaseous products. Similarly, photoexcited electrons can react with O2 to form other reactive oxygen species that can be used to further degrade pollutants [132,137]. The photocatalytic degradation of organic pollutants, such as rhodamine B, imidacloprid, and acetamiprid, has been reported for Ag/AgBr/NiFe2O4 and ZnO/CoFe2O4 [138,139,140,141].
In the case of doped ferrite NPs, the doped element enhances the photocatalytic potential of the pristine ferrite materials; for example, cerium doping in cobalt ferrite NPs plays the important role of electron scavenger [141]. The interfacial electron transfer, photo-excited charge generation, and inhibition of (e-h+) recombination were completely dependent on the Ce 4f level (Equation (1)). The electron-scavenging efficiency of Ce4+ helps to capture the conduction-band electrons, and due to its Lewis acidic nature, it was found to be more effective at trapping electrons compared to the molecular oxygen (Equation (2)). Through an oxidation process (Equation (3)), the electron was transferred to the adsorbed O2, and superoxide radical anions (-O2) were formed [115]. These reactive oxygen species helped in the degradation of the RhB dye.
CoFe2O4 + hv → h+ + e
Ce4+ + e → Ce3+(electron trapping)
Ce3+ + O2 → •O2 + Ce4+(electron transfer)
h+ + H2O → •OH + h+
RhB + ROS → Degraded products
Meanwhile, the H2O molecules can easily scavenge h+ to produce -OH radicals (Equation (4)). The rhodamine B dye can be decomposed into CO2, H2O, and mineral acids.
The incorporation of ferrites into ultraviolet and visible light active metal oxides, polymers, carbon materials, etc. is another way to photodegrade organic pollutants. Figure 7 depicts the transfer of the photogenerated electrons of the metal oxide and ferrites nanocomposites from the conduction band of the one photocatalyst to the other, while holes are jumped from the valence band of the photocatalyst-II to the photocatalyst-I. Reactive species are created as a result of the efficient separation of the electron-hole pair [142,143,144,145,146,147].
The recovery of carbon-based materials is quite difficult due to their poor visible light absorption properties, which decrease their efficiency as photocatalysts for the removal of organic pollutants from wastewater. However, they are viewed as potential resources for improving the efficiency of MFe2O4 as a photocatalyst. Reduced graphene oxide sheets act as an electron trap center, which, as shown in Figure 8a, significantly enhances the photocatalytic activity of the MFe2O4 nanoparticles [148,149,150,151].
Polymers, such as polyaniline, are another type of supporting material. The positions of the highest and lowest occupied molecular orbitals (HOMO and LUMO) are −1.9 and 0.8 eV, respectively, as depicted in Figure 8b. When exposed to visible light, polyaniline on absorption undergoes π-π* transition. When the excited electron in the lowest occupied molecular orbital is transferred to the conduction band, it produces the reactive species, which are responsible for degrading the organic pollutants [152,153].
Ferrite NPs are also combined with a variety of materials, such as metal sulphide [154], metal phosphate [155], metal oxyhalide [156], metal vanadate [3], metal titanate [157], etc., which makes their degradation mechanism different. MgFe2O4/Bi2MoO6 composites have been used to degrade malachite green dye [158]. Free radicals were produced through the Z-scheme mechanism, as shown in Figure 9. Due to the difference in the band gaps of the MgFe2O4 and Bi2MoO6, photogenerated electrons from the conduction band of the latter material diffuse to the valence band and become effectively trapped. Radicals produced by photogenerated holes and electrons, such as hydroxyl and superoxide, are also involved in the photodegradation process. Using the SnFe2O4/ZnFe2O4 heterostructure, the degradation of tetracycline observed followed the S-scheme mechanism (Figure 9), where holes in the valence band of the SnFe2O4 were neutralised by the photogenerated electrons in the ZnFe2O4 [159]. As a result, the availability of holes in the ZnFe2O4 and electrons in the SnFe2O4 makes the production of hydroxyl and superoxide radicals possible.

4.9. Reusability and Stability

The ideal photocatalyst is one that is reusable and has high photocatalytic efficiency under visible light [160]. Therefore, one of the crucial steps is to recycle photocatalysts for the development of heterogeneous photocatalysis. Recent research has shown that magnetic nanoparticles can be successfully recycled using magnetic separation techniques [161,162,163,164,165].
Five consecutive cycles of photodegradation were conducted to examine the stability of Sr0.4Ti0.6Fe2O4.6@SiO2. A slight decrease (<7.0%) in the photocatalytic efficiency of pendimethalin was found after five runs, which was due to adsorption of the intermediate product on the surface of the photocatalyst where active sites are blocked [117]. Similar results were found when the hierarchical nanoflowers of the MgFe2O4, Bentonite and B-, P- co-doped Graphene Oxide were tested for their reusability and stability. The degradation efficiency was found to be decreased by only 5% after 5 consecutive cycles [83].
Generally, most of the magnetic nanocomposites can be reused upto three to four cycles, demonstrating their reusability when used for the photodegradation of pollutants from wastewater [75,101,112,166].

5. Conclusions and Future Perspective

In this review, the impact of different parameters on the photocatalytic degradation of organic effluents using ferrites and their nanocomposites has been elaborated. Different types of photodegradation mechanisms of ferrites and their nanocomposites are discussed. The use of visible light sensitive ferrite-based nanomaterials as photocatalysts for the remediation of wastewater and degradation of organic contaminants like pharmaceuticals, phenols, dyes, and other persistent organic pollutants is one of the most active research areas at the moment. The bulk synthesis of ferrite nanoparticles is needed for practical applications. Using chemical methods, the yield of ferrites is low, whereas physical methods yield ferrites in bulk but in micrometer regime. Future research can be focused on the bulk synthesis of ferrites using green and atomically efficient chemical methods and facile physical methods that can yield ferrites in nanometer range. Moreover, while evaluating the degradation efficacy of a photocatalyst, the pollutant concentration must always be taken into account, regardless of the pollutant that is intended to be removed. So, predicting the pollutant concentration and optimization of the catalyst dose must therefore be taken into account. Generally, ferrites and their nanocomposites, due to their appealing properties, such as their ease of surface modification and stability, are promising nanomaterials for addressing a variety of issues. Even though there have been many studies on ferrites, more research is needed on different parameters that affect the degradation of various organic pollutants under visible light, and technology having the potential to optimize the influential parameters is needed for the photocatalysis process to be more economical.

Author Contributions

Writing—original draft preparation, G.S., M.K., K.J., M.K.U. and C.S.; Writing—review and editing, G.S., M.K., K.J., M.K.U. and C.S.; Supervision, M.K., K.J. and C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anh, H.Q.; Le, T.P.Q.; Da Le, N.; Lu, X.X.; Duong, T.T.; Garnier, J.; Rochelle-Newall, E.; Zhang, S.; Oh, N.-H.; Oeurng, C.; et al. Antibiotics in surface water of East and Southeast Asian countries: A focused review on contamination status, pollution sources, potential risks, and future perspectives. Sci. Total Environ. 2020, 764, 142865. [Google Scholar] [CrossRef] [PubMed]
  2. Show, P.L.; Ooi, C.W.; Anuar, M.S.; Ariff, A.; Yusof, Y.A.; Chen, S.K.; Annuar, M.S.M.; Ling, T.C. Recovery of lipase derived from Burkholderiacenocepacia ST8 using sustainable aqueous two-phase flotation composed of recycling hydrophilic organic solvent and inorganic salt. Sep. Purif. Technol. 2013, 110, 112–118. [Google Scholar] [CrossRef]
  3. Sakhare, P.A.; Pawar, S.S.; Bhat, T.S.; Yadav, S.D.; Patil, G.R.; Patil, P.S.; Sheikh, A.D. Magnetically Recoverable BiVO4/NiFe2O4 Nanocomposite Photocatalyst for Efficient Detoxification of Polluted Water Under Collected Sunlight. Mater. Res. Bull. 2020, 129, 110908. [Google Scholar] [CrossRef]
  4. Saxena, G.; Bharagava, R. Organic and Inorganic Pollutants in Industrial Wastes: Ecotoxicological Effects, Health Hazards, and Bioremediation Approaches. In Environmental Pollutants and Their Bioremediation Approaches; Bharagava, R., Ed.; CRC Press: Boca Raton, FL, USA, 2017; pp. 23–56. ISBN 9781138628892. [Google Scholar]
  5. Uday, U.; Mahata, N.; Sasmal, S.; Bandyopadhyay, T.; Mondal, A.; Bhunia, B. Dyes Contamination in the Environment: Ecotoxicological Effects, Health Hazards, and Biodegradation and Bioremediation Mechanisms for Environmental Cleanup. In Environmental Pollutants and Their Bioremediation Approaches; Bharagava, R., Ed.; CRC Press: Boca Raton, FL, USA, 2017; pp. 127–176. ISBN 9781138628892. [Google Scholar]
  6. Koduru, J.; Shankar, S.; More, N.; Shikha; Lingamdinne, L.; Singh, J. Toxic Metals Contamination in the Environment: Toxi-cological Effects and Bioremediation Approaches for Environmental Cleanup. In Environmental Pollutants and Their Bioremediation Approaches; Bharagava, R., Ed.; CRC Press: Boca Raton, FL, USA, 2017; pp. 209–240. ISBN 9781138628892. [Google Scholar]
  7. Karimi-Maleh, H.; Alizadeh, M.; Orooji, Y.; Karimi, F.; Baghayeri, M.; Rouhi, J.; Tajik, S.; Beitollahi, H.; Agarwal, S.; Gupta, V.K.; et al. Guanine-Based DNA Biosensor Amplified with Pt/SWCNTs Nanocomposite as Analytical Tool for Nanomolar Determination of Daunorubicin as an Anticancer Drug: A Docking/Experimental Investigation. Ind. Eng. Chem. Res. 2021, 60, 816–823. [Google Scholar] [CrossRef]
  8. Deliyanni, E.A.; Kyzas, G.Z.; Matis, K.A. Various flotation techniques for metal ions removal. J. Mol. Liq. 2016, 225, 260–264. [Google Scholar] [CrossRef]
  9. Yu, K.L.; Show, P.L.; Ong, H.C.; Ling, T.C.; Lan, J.C.-W.; Chen, W.-H.; Chang, J.-S. Microalgae from wastewater treatment to biochar—Feedstock preparation and conversion technologies. Energy Convers. Manag. 2017, 150, 1–13. [Google Scholar] [CrossRef]
  10. Chua, S.F.; Nouri, A.; Ang, W.L.; Mahmoudi, E.; Mohammad, A.W.; Benamor, A.; Ba-Abbad, M. The emergence of multifunctional adsorbents and their role in environmental remediation. J. Environ. Chem. Eng. 2021, 9, 104793. [Google Scholar] [CrossRef]
  11. Liu, Y.; Deng, Y.-Y.; Zhang, Q.; Liu, H. Overview of recent developments of resource recovery from wastewater via electrochemistry-based technologies. Sci. Total Environ. 2021, 757, 143901. [Google Scholar] [CrossRef]
  12. Hansima, M.A.C.K.; Makehelwala, M.; Jinadasa, K.B.S.N.; Wei, Y.; Nanayakkara, K.G.N.; Herath, A.C.; Weerasooriya, R. Fouling of ion exchange membranes used in the electrodialysis reversal advanced water treatment: A review. Chemosphere 2021, 263, 127951. [Google Scholar] [CrossRef]
  13. Ajiboye, T.O.; Oyewo, O.A.; Onwudiwe, D.C. Simultaneous removal of organics and heavy metals from industrial wastewater: A review. Chemosphere 2021, 262, 128379. [Google Scholar] [CrossRef]
  14. Arshad, Z.I.M.; Amid, A.; Yusof, F.; Jaswir, I.; Ahmad, K.; Loke, S.P. Bromelain: An overview of industrial application and purification strategies. Appl. Microbiol. Biotechnol. 2014, 98, 7283–7297. [Google Scholar] [CrossRef] [Green Version]
  15. Naushad, M.; Sharma, G.; Alothman, Z.A. Photodegradation of toxic dye using Gum Arabic-crosslinked-poly(acrylamide)/Ni(OH)2/FeOOH nanocomposites hydrogel. J. Clean. Prod. 2019, 241, 118263. [Google Scholar] [CrossRef]
  16. Faisal, A.A.H.; Al-Wakel, S.F.A.; Assi, H.A.; Naji, L.A.; Naushad, M. Waterworks sludge-filter sand permeable reactive barrier for removal of toxic lead ions from contaminated groundwater. J. Water Process Eng. 2020, 33, 101112. [Google Scholar] [CrossRef]
  17. Mironyuk, I.; Tatarchuk, T.; Naushad, M.; Vasylyeva, H.; Mykytyn, I. Highly efficient adsorption of strontium ions by carbonated mesoporous TiO2. J. Mol. Liq. 2019, 285, 742–753. [Google Scholar] [CrossRef]
  18. Chen, C.; Duan, F.; Zhao, S.; Wang, W.; Yang, F.; Nuansing, W.; Zhang, B.; Qin, Y.; Knez, M. Porous Fe2O3 nanotubes with α-γ phase junction for enhanced charge separation and photocatalytic property produced by molecular layer deposition. Appl. Catal. B Environ. 2019, 248, 218–225. [Google Scholar] [CrossRef]
  19. Crini, G.; Lichtfouse, E. Advantages and disadvantages of techniques used for wastewater treatment. Environ. Chem. Lett. 2019, 17, 145–155. [Google Scholar] [CrossRef]
  20. Ong, C.B.; Ng, L.Y.; Mohammad, A.W. A review of ZnO nanoparticles as solar photocatalysts: Synthesis, mechanisms and applications. Renew. Sustain. Energy Rev. 2018, 81, 536–551. [Google Scholar] [CrossRef]
  21. Khalid, N.R.; Majid, A.; Tahir, M.B.; Niaz, N.A.; Khalid, S. Carbonaceous-TiO2 nanomaterials for photocatalytic degradation of pollutants: A review. Ceram. Int. 2017, 43, 14552–14571. [Google Scholar] [CrossRef]
  22. Jiang, M.; Zhang, M.; Wang, L.; Fei, Y.; Wang, S.; Núñez-Delgado, A.; Bokhari, A.; Race, M.; Khataee, A.; Klemeš, J.J.; et al. Photocatalytic degradation of xanthate in flotation plant tailings by TiO2/graphene nanocomposites. Chem. Eng. J. 2022, 431, 134104. [Google Scholar] [CrossRef]
  23. Hitam, C.N.C.; Jalil, A.A.; Abdulrasheed, A.A. A review on recent progression of photocatalytic desulphurization study over decorated photocatalysts. J. Ind. Eng. Chem. 2019, 74, 172–186. [Google Scholar] [CrossRef]
  24. Byrne, C.; Subramanian, G.; Pillai, S.C. Recent advances in photocatalysis for environmental applications. J. Environ. Chem. Eng. 2018, 6, 3531–3555. [Google Scholar] [CrossRef]
  25. Khaki, M.R.D.; Shafeeyan, M.S.; Raman, A.A.A.; Daud, W.M.A.W. Application of doped photocatalysts for organic pollutant degradation—A review. J. Environ. Manag. 2017, 198, 78–94. [Google Scholar] [CrossRef]
  26. Danish, M.S.S.; Estrella, L.L.; Alemaida, I.M.A.; Lisin, A.; Moiseev, N.; Ahmadi, M.; Nazari, M.; Wali, M.; Zaheb, H.; Senjyu, T. Photocatalytic Applications of Metal Oxides for Sustainable Environmental Remediation. Metals 2021, 11, 80. [Google Scholar] [CrossRef]
  27. Chakkunny, A.H.; Thomas, B.; Alexander, L.K. Influence of pH and salinity on the photocatalytic dye degradation and heavy metal ion reduction with cobalt ferrite photocatalysts. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1187, 012018. [Google Scholar] [CrossRef]
  28. Helmy, E.T.; Soliman, U.A.; Elbasiony, A.M.; Nguyen, B.-S. CuCe-Ferrite/TiO2 Nanocomposite as an Efficient Magnetically Separable Photocatalyst for Dye Pollutants Decolorization. Top. Catal. 2023, 66, 53–63. [Google Scholar] [CrossRef]
  29. Padmapriya, G.; Manikandan, A.; Krishnasamy, V.; Jaganathan, S.K.; Antony, S.A. Spinel Ni Zn1Fe2O4 (0.0 ≤ x ≤ 1.0) nano-photocatalysts: Synthesis, characterization and photocatalytic degradation of methylene blue dye. J. Mol. Struct. 2016, 1119, 39–47. [Google Scholar] [CrossRef]
  30. Haruna, A.; Abdulkadir, I.; Idris, S.O. Photocatalytic activity and doping effects of BiFeO3 nanoparticles in model organic dyes. Heliyon 2020, 6, e03237. [Google Scholar] [CrossRef] [Green Version]
  31. Oliveira, T.P.; Marques, G.N.; Castro, M.A.M.; Costa, R.C.V.; Rangel, J.H.G.; Rodrigues, S.F.; dos Santos, C.C.; Oliveira, M.M. Synthesis and photocatalytic investigation of ZnFe2O4 in the degradation of organic dyes under visible light. J. Mater. Res. Technol. 2020, 9, 15001–15015. [Google Scholar] [CrossRef]
  32. Hitam, C.; Jalil, A. A review on exploration of Fe2O3 photocatalyst towards degradation of dyes and organic contaminants. J. Environ. Manag. 2020, 258, 110050. [Google Scholar] [CrossRef]
  33. Ji, J.; Huang, Y.; Yin, J.; Zhao, X.; Cheng, X.; He, S.; Li, X.; He, J.; Liu, J. Synthesis and Electromagnetic and Microwave Absorption Properties of Monodispersive Fe3O4/α-Fe2O3 Composites. ACS Appl. Nano Mater. 2018, 1, 3935–3944. [Google Scholar] [CrossRef]
  34. Saif, S.; Tahir, A.; Asim, T.; Chen, Y.; Adil, S.F. Polymeric Nanocomposites of Iron–Oxide Nanoparticles (IONPs) Synthesized Using Terminalia chebula Leaf Extract for Enhanced Adsorption of Arsenic(V) from Water. Colloids Interfaces 2019, 3, 17. [Google Scholar] [CrossRef] [Green Version]
  35. Arias, L.S.; Pessan, J.P.; Vieira, A.P.M.; de Lima, T.M.T.; Delbem, A.C.B.; Monteiro, D.R. Iron Oxide Nanoparticles for Biomedical Applications: A Perspective on Synthesis, Drugs, Antimicrobial Activity, and Toxicity. Antibiotics 2018, 7, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Dillert, R.; Taffa, D.H.; Wark, M.; Bredow, T.; Bahnemann, D.W. Research Update: Photoelectrochemical water splitting and photocatalytic hydrogen production using ferrites (MFe2O4) under visible light irradiation. APL Mater. 2015, 3, 104001. [Google Scholar] [CrossRef] [Green Version]
  37. Casbeer, E.; Sharma, V.K.; Li, X.-Z. Synthesis and photocatalytic activity of ferrites under visible light: A review. Sep. Purif. Technol. 2012, 87, 1–14. [Google Scholar] [CrossRef]
  38. Buchanan, R.C. (Ed.) Ceramic Materials for Electronics, 3rd ed.; Dekker: New York, NY, USA; Basel, Switzerland, 2004; revised expanded; ISBN 9780824740283. [Google Scholar]
  39. Tokubuchi, T.; Arbi, R.I.; Zhenhua, P.; Katayama, K.; Turak, A.; Sohn, W.Y. Enhanced photoelectrochemical water splitting efficiency of hematite (α-Fe2O3)-Based photoelectrode by the introduction of maghemite (γ-Fe2O3) nanoparticles. J. Photochem. Photobiol. A Chem. 2021, 410, 113179. [Google Scholar] [CrossRef]
  40. Kefeni, K.K.; Mamba, B.B. Photocatalytic application of spinel ferrite nanoparticles and nanocomposites in wastewater treatment: Review. Sustain. Mater. Technol. 2020, 23, e00140. [Google Scholar] [CrossRef]
  41. Kushwaha, P.; Chauhan, P. Synthesis of spherical and Rod-Like EDTA assisted α-Fe2O3 nanoparticles via Co-precipitation method. Mater. Today Proc. 2021, 44, 3086–3090. [Google Scholar] [CrossRef]
  42. Uma, K.; KrishnaKumar, B.; Pan, G.-T.; Yang, T.C.-K.; Lin, J.-H. Enriched silver plasmon resonance activity on the sonochemical synthesis of ZnO flowers with α-Fe2O3 as an efficient catalyst for photo-Fenton reaction and photo-oxidation of ethanol. J. Water Process Eng. 2020, 34, 101089. [Google Scholar] [CrossRef]
  43. Li, Y.; Huang, Y.; Zheng, Y.; Huang, R.; Yao, J. Facile and efficient synthesis of α-Fe2O3 nanocrystals by glucose-assisted thermal decomposition method and its application in lithium ion batteries. J. Power Sour. 2019, 416, 62–71. [Google Scholar] [CrossRef]
  44. Tadic, M.; Panjan, M.; Tadic, B.V.; Lazovic, J.; Damnjanovic, V.; Kopani, M.; Kopanja, L. Magnetic properties of hematite (α − Fe2O3) nanoparticles synthesized by sol-gel synthesis method: The influence of particle size and particle size distribution. J. Electr. Eng. 2019, 70, 71–76. [Google Scholar] [CrossRef] [Green Version]
  45. Luo, S.-H.; Hu, D.-B.; Liu, H.; Li, J.-Z.; Yi, T.-F. Hydrothermal synthesis and characterization of α-Fe2O3/C using acid-pickled iron oxide red for Li-ion batteries. J. Hazard. Mater. 2019, 368, 714–721. [Google Scholar] [CrossRef] [PubMed]
  46. Bozkurt, G. Synthesis and Characterization of α-Fe2O3 Nanoparticles by Microemulsion Method. Erzincan Üniv. Fen Bilim. Enst. Derg. 2020, 13, 890–897. [Google Scholar] [CrossRef]
  47. He, H.-Y. Photocatalytic degradations of dyes on magnetically separable Ni1−xCoxFe2O4 nanoparticles synthesized by a hydrothermal process. Part. Sci. Technol. 2016, 34, 143–151. [Google Scholar] [CrossRef]
  48. Parsons, J.G.; Lopez, M.L.; Peralta-Videa, J.R.; Gardea-Torresdey, J.L. Determination of arsenic (III) and arsenic (V) binding to microwave assisted hydrothermal synthetically prepared Fe3O4, Mn3O4, and MnFe2O4 nanoadsorbents. Microchem. J. 2009, 91, 100–106. [Google Scholar] [CrossRef]
  49. Kaur, N.; Kaur, M. Comparative studies on impact of synthesis methods on structural and magnetic properties of magnesium ferrite nanoparticles. Process. Appl. Ceram. 2014, 8, 137–143. [Google Scholar] [CrossRef] [Green Version]
  50. Routray, K.L.; Saha, S.; Behera, D. Green synthesis approach for nano sized CoFe2O4 through aloe vera mediated sol-gel auto combustion method for high frequency devices. Mater. Chem. Phys. 2019, 224, 29–35. [Google Scholar] [CrossRef]
  51. Zhang, D.; Pu, X.; Du, K.; Yu, Y.M.; Shim, J.J.; Cai, P.; Kim, S.I.; Seo, H.J. Combustion synthesis of magnetic Ag/NiFe2O4 composites with enhanced visible-light photocatalytic properties. Sep. Purif. Technol. 2014, 137, 82–85. [Google Scholar] [CrossRef]
  52. Johnson, B.F.G. Nanoparticles in Catalysis. Top. Catal. 2003, 24, 147–159. [Google Scholar] [CrossRef]
  53. Govan, J.; Gun’Ko, Y.K. Recent Advances in the Application of Magnetic Nanoparticles as a Support for Homogeneous Catalysts. Nanomaterials 2014, 4, 222–241. [Google Scholar] [CrossRef] [Green Version]
  54. Demirelli, M.; Karaoglu, E.; Baykal, A.; Sozeri, H. M-hexaferrite–APTES/Pd(0) Magnetically Recyclable Nano Catalysts (MRCs). J. Inorg. Organomet. Polym. Mater. 2013, 23, 1274–1281. [Google Scholar] [CrossRef]
  55. Demirelli, M.; Karaoğlu, E.; Baykal, A.; Sözeri, H.; Uysal, E.; Duygulu, O. Recyclable NiFe2O4–APTES/Pd Magnetic Nanocatalyst. J. Inorg. Organomet. Polym. Mater. 2013, 23, 937–943. [Google Scholar] [CrossRef]
  56. Gawande, M.B.; Branco, P.S.; Nogueira, I.D.; Ghumman, C.A.A.; Bundaleski, N.; Santos, A.; Teodoro, O.M.N.D.; Luque, R. Catalytic applications of a versatile magnetically separable Fe–Mo (Nanocat-Fe–Mo) nanocatalyst. Green Chem. 2013, 15, 682–689. [Google Scholar] [CrossRef]
  57. Gawande, M.B.; Rathi, A.K.; Branco, P.S.; Nogueira, I.D.; Velhinho, A.; Shrikhande, J.J.; Indulkar, U.U.; Jayaram, R.V.; Ghumman, C.A.A.; Bundaleski, N.; et al. Regio- and Chemoselective Reduction of Nitroarenes and Carbonyl Compounds over Recyclable Magnetic Ferrite-Nickel Nanoparticles (Fe3O4-Ni) by Using Glycerol as a Hydrogen Source. Chem. Eur. J. 2012, 18, 12628–12632. [Google Scholar] [CrossRef] [PubMed]
  58. Gawande, M.B.; Velhinho, A.; Nogueira, I.D.; Ghumman, C.A.A.; Teodoro, O.M.N.D.; Branco, P.S. A facile synthesis of cysteine–ferrite magnetic nanoparticles for application in multicomponent reactions—A sustainable protocol. RSC Adv. 2012, 2, 6144–6149. [Google Scholar] [CrossRef]
  59. Gawande, M.B.; Rathi, A.; Nogueira, I.D.; Ghumman, C.A.A.; Bundaleski, N.; Teodoro, O.M.N.D.; Branco, P.S. A Recyclable Ferrite-Co Magnetic Nanocatalyst for the Oxidation of Alcohols to Carbonyl Compounds. Chempluschem 2012, 77, 865–871. [Google Scholar] [CrossRef]
  60. Duan, H.-Z.; Zhou, F.-L.; Cheng, X.; Chen, G.-H.; Li, Q.-L. Preparation of hollow microspheres of Ce3+ dopedNiCo ferrite with high microwave absorbing performance. J. Magn. Magn. Mater. 2017, 424, 467–471. [Google Scholar] [CrossRef]
  61. Gan, L.; Shang, S.; Yuen, C.W.M.; Jiang, S.-X.; Hu, E. Hydrothermal synthesis of magnetic CoFe2O4/graphene nanocomposites with improved photocatalytic activity. Appl. Surf. Sci. 2015, 351, 140–147. [Google Scholar] [CrossRef]
  62. Rehman, M.A.; Yusoff, I.; Alias, Y. Structural, morphological and magnetic investigations of CuCe0.2Fe1.8O4 graphene-supported nanocomposites. Ceram. Int. 2016, 42, 1399–1407. [Google Scholar] [CrossRef]
  63. Wang, B.; Si, L.; Geng, J.; Su, Y.; Li, Y.; Yan, X.; Chen, L. Controllable magnetic 3D nitrogen-doped graphene gel: Synthesis, characterization, and catalytic performance. Appl. Catal. B Environ. 2017, 204, 316–323. [Google Scholar] [CrossRef]
  64. Randhawa, B.S.; Dosanjh, H.S.; Kaur, M. Preparation of Ferrites from the Combustion of Metal Nitrate-Oxalyl Dihydrazide Solutions. IJEMS 2005, 12, 151–154. [Google Scholar]
  65. Li, X.; Hou, Y.; Zhao, Q.; Wang, L. A general, one-step and template-free synthesis of sphere-like zinc ferrite nanostructures with enhanced photocatalytic activity for dye degradation. J. Colloid Interface Sci. 2011, 358, 102–108. [Google Scholar] [CrossRef]
  66. Dom, R.; Subasri, R.; Radha, K.; Borse, P.H. Synthesis of solar active nanocrystalline ferrite, MFe2O4 (M: Ca, Zn, Mg) photocatalyst by microwave irradiation. Solid State Commun. 2011, 151, 470–473. [Google Scholar] [CrossRef]
  67. Borhan, A.I.; Samoila, P.; Hulea, V.; Iordan, A.R.; Palamaru, M.N. Photocatalytic activity of spinel ZnFe2−xCrxO4 nanoparticles on removal Orange I azo dye from aqueous solution. J. Taiwan Inst. Chem. Eng. 2014, 45, 1655–1660. [Google Scholar] [CrossRef]
  68. Borhan, A.I.; Samoila, P.; Hulea, V.; Iordan, A.R.; Palamaru, M.N. Effect of Al3+ substituted zinc ferrite on photocatalytic degradation of Orange I azo dye. J. Photochem. Photobiol. A Chem. 2014, 279, 17–23. [Google Scholar] [CrossRef]
  69. Jesudoss, S.K.; Vijaya, J.J.; Kennedy, L.J.; Rajan, P.I.; Al-Lohedan, H.A.; Ramalingam, R.J.; Kaviyarasu, K.; Bououdina, M. Studies on the efficient dual performance of Mn1–xNixFe2O4 spinel nanoparticles in photodegradation and antibacterial activity. J. Photochem. Photobiol. B Biol. 2016, 165, 121–132. [Google Scholar] [CrossRef]
  70. Zhang, Z.; Wang, W. Solution combustion synthesis of CaFe2O4 nanocrystal as a magnetically separable photocatalyst. Mater. Lett. 2014, 133, 212–215. [Google Scholar] [CrossRef]
  71. Shahid, M.; Jingling, L.; Ali, Z.; Shakir, I.; Warsi, M.F.; Parveen, R.; Nadeem, M. Photocatalytic degradation of methylene blue on magnetically separable MgFe2O4 under visible light irradiation. Mater. Chem. Phys. 2013, 139, 566–571. [Google Scholar] [CrossRef]
  72. Zhu, Z.; Li, X.; Zhao, Q.; Li, Y.; Sun, C.; Cao, Y. Photocatalytic performances and activities of Ag-doped CuFe2O4 nanoparticles. Mater. Res. Bull. 2013, 48, 2927–2932. [Google Scholar] [CrossRef]
  73. Shetty, K.; Renuka, L.; Nagaswarupa, H.P.; Nagabhushana, H.; Anantharaju, K.S.; Rangappa, D.; Prashantha, S.C.; Ashwini, K. A comparative study on CuFe2O4, ZnFe2O4 and NiFe2O4: Morphology, Impedance and Photocatalytic studies. Mater. Today Proc. 2017, 4, 11806–11815. [Google Scholar] [CrossRef]
  74. Lassoued, A.; Lassoued, M.S.; Dkhil, B.; Ammar, S.; Gadri, A. RETRACTED ARTICLE: Photocatalytic degradation of methyl orange dye by NiFe2O4 nanoparticles under visible irradiation: Effect of varying the synthesis temperature. J. Mater. Sci. Mater. Electron. 2018, 29, 7057–7067. [Google Scholar] [CrossRef]
  75. Roonasi, P.; Mazinani, M. Synthesis and application of barium ferrite/activated carbon composite as an effective solar photocatalyst for discoloration of organic dye contaminants in wastewater. J. Environ. Chem. Eng. 2017, 5, 3822–3827. [Google Scholar] [CrossRef]
  76. Talebi, R. Preparation and characterization of cobalt ferrite nanoparticles with different capping agents and its photocatalyst application. J. Mater. Sci. Mater. Electron. 2017, 28, 9749–9754. [Google Scholar] [CrossRef]
  77. Talukdar, S.; Mandal, D.; Mandal, K. Surface modification of Cobalt ferrite nano-hollowspheres for inherent multiple photoluminescence and enhanced photocatalytic activities. Chem. Phys. Lett. 2017, 672, 57–62. [Google Scholar] [CrossRef]
  78. Parhizkar, J.; Habibi, M.H.; Mosavian, S.Y. Synthesis and Characterization of Nano CoFe2O4 Prepared by Sol-Gel Auto-Combustion with Ultrasonic Irradiation and Evaluation of Photocatalytic Removal and Degradation Kinetic of Reactive Red 195. Silicon 2019, 11, 1119–1129. [Google Scholar] [CrossRef]
  79. Fang, M.-M.; Shao, J.-X.; Huang, X.-G.; Wang, J.-Y.; Chen, W. Direct Z-scheme CdFe2O4/g-C3N4 hybrid photocatalysts for highly efficient ceftiofur sodium photodegradation. J. Mater. Sci. Technol. 2020, 56, 133–142. [Google Scholar] [CrossRef]
  80. Bo, L.; Hu, Y.; Zhang, Z.; Tong, J. Efficient photocatalytic degradation of Rhodamine B catalyzed by SrFe2O4/g-C3N4 composite under visible light. Polyhedron 2019, 168, 94–100. [Google Scholar] [CrossRef]
  81. Shakir, I.; Sarfraz, M.; Ali, Z.; Aboud, M.F.A.; Agboola, P.O. Magnetically separable and recyclable graphene-MgFe2O4 nanocomposites for enhanced photocatalytic applications. J. Alloys Compd. 2016, 660, 450–455. [Google Scholar] [CrossRef]
  82. Grewal, J.K.; Kaur, M.; Sharma, R.K.; Oliveira, A.C.; Garg, V.K.; Sharma, V.K. Structural and Photocatalytic Studies on Oxygen Hyperstoichiometric Titanium-Substituted Strontium Ferrite Nanoparticles. Magnetochemistry 2022, 8, 120. [Google Scholar] [CrossRef]
  83. Ubhi, M.K.; Kaur, M.; Singh, D.; Javed, M.; Oliveira, A.C.; Garg, V.K.; Sharma, V.K. Hierarchical Nanoflowers of MgFe2O4, Bentonite and B-,P- Co-Doped Graphene Oxide as Adsorbent and Photocatalyst: Optimization of Parameters by Box–Behnken Methodology. Int. J. Mol. Sci. 2022, 23, 9678. [Google Scholar] [CrossRef]
  84. Singh, E.; Kaur, M.; Sharma, S. Structural tuning of CTAB@MgFe2O4 nanocomposite as peroxidase mimic for H2O2 and glucose sensing. Mater. Chem. Phys. 2021, 271, 124851. [Google Scholar] [CrossRef]
  85. Kaur, M.; Kaur, M.; Singh, D.; Oliveira, A.C.; Garg, V.K.; Sharma, V.K. Synthesis of CaFe2O4-NGO Nanocomposite for Effective Removal of Heavy Metal Ion and Photocatalytic Degradation of Organic Pollutants. Nanomaterials 2021, 11, 1471. [Google Scholar] [CrossRef] [PubMed]
  86. Kaur, M.; Ubhi, M.K.; Grewal, J.K.; Singh, D. Insight into the structural, optical, adsorptive, and photocatalytic properties of MgFe2O4-bentonite nanocomposites. J. Phys. Chem. Solids 2021, 154, 110060. [Google Scholar] [CrossRef]
  87. Khushboo; Kaur, M.; Jeet, K. Mechanistic insight into adsorption and photocatalytic potential of magnesium ferrite-bentonite nanocomposite. J. Photochem. Photobiol. A Chem. 2022, 425, 113717. [Google Scholar] [CrossRef]
  88. Ubhi, M.K.; Kaur, M.; Grewal, J.K.; Oliveira, A.C.; Garg, V.K.; Sharma, V.K. Insight into photocatalytic behavior of magnesium ferrite–bentonite nanocomposite for the degradation of organic contaminants. J. Mater. Res. 2022, 38, 990–1006. [Google Scholar] [CrossRef]
  89. Guan, X.; Kuang, J.; Yang, L.; Lu, M.; Wang, G. Membrane-Solvothermal Synthesis of Cobalt Ferrite/Reduced Graphene Oxide Nanocomposites and Their Photocatalytic and Electromagnetic Wave Absorption Properties. ChemistrySelect 2019, 4, 9516–9522. [Google Scholar] [CrossRef]
  90. Sheshmani, S.; Falahat, B.; Nikmaram, F.R. Preparation of magnetic graphene oxide-ferrite nanocomposites for oxidative decomposition of Remazol Black B. Int. J. Biol. Macromol. 2017, 97, 671–678. [Google Scholar] [CrossRef]
  91. Yoo, P.S.; Reddy, D.A.; Jia, Y.; Bae, S.E.; Huh, S.; Liu, C. Magnetic core-shell ZnFe2O4/ZnS nanocomposites for photocatalytic application under visible light. J. Colloid Interface Sci. 2017, 486, 136–143. [Google Scholar] [CrossRef]
  92. Middea, A.; Spinelli, L.S.; Jr, F.G.S.; Neumann, R.; Fernandes, T.L.A.P.; Gomes, O.D.F.M. Preparation and characterization of an organo-palygorskite-Fe3O4 nanomaterial for removal of anionic dyes from wastewater. Appl. Clay Sci. 2017, 139, 45–53. [Google Scholar] [CrossRef]
  93. Song, W.; Gao, B.; Xu, X.; Xing, L.; Han, S.; Duan, P.; Song, W.; Jia, R. Adsorption–desorption behavior of magnetic amine/Fe3O4 functionalized biopolymer resin towards anionic dyes from wastewater. Bioresour. Technol. 2016, 210, 123–130. [Google Scholar] [CrossRef]
  94. Soltani, T.; Entezari, M.H. Photolysis and photocatalysis of methylene blue by ferrite bismuth nanoparticles under sunlight irradiation. J. Mol. Catal. A Chem. 2013, 377, 197–203. [Google Scholar] [CrossRef]
  95. Bian, Z.; Huo, Y.; Zhang, Y.; Zhu, J.; Lu, Y.; Li, H. Aerosol-spay assisted assembly of Bi2Ti2O7 crystals in uniform porous microspheres with enhanced photocatalytic activity. Appl. Catal. B Environ. 2009, 91, 247–253. [Google Scholar] [CrossRef]
  96. Harraz, F.A.; Mohamed, R.M.; Rashad, M.M.; Wang, Y.C.; Sigmund, W. Magnetic nanocomposite based on titania–silica/cobalt ferrite for photocatalytic degradation of methylene blue dye. Ceram. Int. 2014, 40, 375–384. [Google Scholar] [CrossRef]
  97. Shih, Y.-J.; Su, C.-C.; Chen, C.-W.; Dong, C.-D. Synthesis of magnetically recoverable ferrite (MFe2O4, M Co, Ni and Fe)-supported TiO2 photocatalysts for decolorization of methylene blue. Catal. Commun. 2015, 72, 127–132. [Google Scholar] [CrossRef]
  98. Nardea, S.; Gadegoneb, S.; Lanjewar, M.R.; Lanjewar, R.B. Degradation of Azo Dye Congo Red Using Ni0.6Co0.4Fe2O4 as Photocatalyst. Pharma Chem. 2017, 9, 115–120. [Google Scholar]
  99. Ge, M.; Liu, W.; Hu, X.-R.; Li, Z.-L. Magnetically separable Ag/AgBr/NiFe2O4 composite as a highly efficient visible light plasmonic photocatalyst. J. Phys. Chem. Solids 2017, 109, 1–8. [Google Scholar] [CrossRef]
  100. Abdel Maksoud, M.I.A.; El-Sayyad, G.S.; El-Khawaga, A.M.; Abd Elkodous, M.; Abokhadra, A.; Elsayed, M.A.; Gobara, M.; Soliman, L.I.; El-Bahnasawy, H.H.; Ashour, A.H. Nanostructured Mg substituted Mn-Zn ferrites: A magnetic recyclable catalyst for outstanding photocatalytic and antimicrobial potentials. J. Hazard. Mater. 2020, 399, 123000. [Google Scholar] [CrossRef]
  101. Singh, S.; Kaur, P.; Bansal, S.; Singhal, S. Enhanced photocatalytic performance of Ru-doped spinel nanoferrites for treating recalcitrant organic pollutants in wastewater. J. Sol-Gel Sci. Technol. 2019, 92, 760–774. [Google Scholar] [CrossRef]
  102. Tabasum, A.; Alghuthaymi, M.; Qazi, U.Y.; Shahid, I.; Abbas, Q.; Javaid, R.; Nadeem, N.; Zahid, M. UV-Accelerated Photocatalytic Degradation of Pesticide over Magnetite and Cobalt Ferrite Decorated Graphene Oxide Composite. Plants 2020, 10, 6. [Google Scholar] [CrossRef] [PubMed]
  103. Naghizadeh, M.; Taher, M.A.; Tamaddon, A.-M. Facile synthesis and characterization of magnetic nanocomposite ZnO/CoFe2O4 hetero-structure for rapid photocatalytic degradation of imidacloprid. Heliyon 2019, 5, e02870. [Google Scholar] [CrossRef] [Green Version]
  104. Wang, M.; Fang, G.; Liu, P.; Zhou, D.; Ma, C.; Zhang, D.; Zhan, J. Fe3O4@β-CD nanocomposite as heterogeneous Fenton-like catalyst for enhanced degradation of 4-chlorophenol (4-CP). Appl. Catal. B Environ. 2016, 188, 113–122. [Google Scholar] [CrossRef]
  105. Boruah, P.K.; Sharma, B.; Karbhal, I.; Shelke, M.V.; Das, M.R. Ammonia-modified graphene sheets decorated with magnetic Fe3O4 nanoparticles for the photocatalytic and photo-Fenton degradation of phenolic compounds under sunlight irradiation. J. Hazard. Mater. 2016, 325, 90–100. [Google Scholar] [CrossRef] [PubMed]
  106. Hou, Y.; Li, X.; Zhao, Q.; Chen, G. ZnFe2O4 multi-porous microbricks/graphene hybrid photocatalyst: Facile synthesis, improved activity and photocatalytic mechanism. Appl. Catal. B Environ. 2013, 142–143, 80–88. [Google Scholar] [CrossRef]
  107. Yang, J.; Chen, H.; Gao, J.; Yan, T.; Zhou, F.; Cui, S.; Bi, W. Synthesis of Fe3O4/g-C3N4 nanocomposites and their application in the photodegradation of 2,4,6-trichlorophenol under visible light. Mater. Lett. 2016, 164, 183–189. [Google Scholar] [CrossRef]
  108. Stoia, M.; Muntean, C.; Militaru, B. MnFe2O4 nanoparticles as new catalyst for oxidative degradation of phenol by peroxydisulfate. J. Environ. Sci. 2017, 53, 269–277. [Google Scholar] [CrossRef]
  109. Arifin, N.; Karim, K.M.R.; Abdullah, H.; Khan, M.R. Synthesis of Titania Doped Copper Ferrite Photocatalyst and Its Photoactivity towards Methylene Blue Degradation under Visible Light Irradiation. Bull. Chem. React. Eng. Catal. 2019, 14, 219–227. [Google Scholar] [CrossRef] [Green Version]
  110. Mapossa, A.B.; Mhike, W.; Adalima, J.L.; Tichapondwa, S. Removal of Organic Dyes from Water and Wastewater Using Magnetic Ferrite-Based Titanium Oxide and Zinc Oxide Nanocomposites: A Review. Catalysts 2021, 11, 1543. [Google Scholar] [CrossRef]
  111. Gupta, N.K.; Ghaffari, Y.; Kim, S.; Bae, J.; Kim, K.S.; Saifuddin, M. Photocatalytic Degradation of Organic Pollutants over MFe2O4 (M = Co, Ni, Cu, Zn) Nanoparticles at Neutral PH. Sci. Rep. 2020, 10, 4942. [Google Scholar] [CrossRef] [Green Version]
  112. Makofane, A.; Motaung, D.E.; Hintsho-Mbita, N.C. Photocatalytic degradation of methylene blue and sulfisoxazole from water using biosynthesized zinc ferrite nanoparticles. Ceram. Int. 2021, 47, 22615–22626. [Google Scholar] [CrossRef]
  113. Rashmi, S.K.; Bhojya Naik, H.S.; Jayadevappa, H.; Viswanath, R.; Patil, S.B.; Madhukara Naik, M. Solar light responsive Sm-Zn ferrite nanoparticle as efficient photocatalyst. Mater. Sci. Eng. B 2017, 225, 86–97. [Google Scholar] [CrossRef]
  114. Kirankumar, V.S.; Sumathi, S. Copper and cerium co-doped cobalt ferrite nanoparticles: Structural, morphological, optical, magnetic, and photocatalytic properties. Environ. Sci. Pollut. Res. 2019, 26, 19189–19206. [Google Scholar] [CrossRef]
  115. Mahdikhah, V.; Saadatkia, S.; Sheibani, S.; Ataie, A. Outstanding photocatalytic activity of CoFe2O4/rGO nanocomposite in degradation of organic dyes. Opt. Mater. 2020, 108, 110193. [Google Scholar] [CrossRef]
  116. Manikandan, A.; Hema, E.; Durka, M.; Seevakan, K.; Alagesan, T.; Arul Antony, S. Room Temperature Ferromagnetism of Magnetically Recyclable Photocatalyst of Cu1−xMnxFe2O4-TiO2 (0.0 ≤ x ≤ 0.5) Nanocomposites. J. Supercond. Nov. Magn. 2015, 28, 1783–1795. [Google Scholar] [CrossRef]
  117. Grewal, J.K.; Kaur, M.; Ubhi, M.K.; Oliveira, A.C.; Garg, V.K.; Sharma, V.K. Structural, magnetic, and photocatalytic properties of core–shell reversal nanocomposites of titanium-doped strontium ferrite and silica. J. Mater. Res. 2023, 38, 1019–1034. [Google Scholar] [CrossRef]
  118. Mohan, H.; Ramalingam, V.; Adithan, A.; Natesan, K.; Seralathan, K.-K.; Shin, T. Highly efficient visible light driven photocatalytic activity of zinc/ferrite: Carbamazepine degradation, mechanism and toxicity assessment. J. Hazard. Mater. 2021, 416, 126209. [Google Scholar] [CrossRef] [PubMed]
  119. Nguyen, T.B.; Huang, C.P.; Doong, R.-A. Photocatalytic degradation of bisphenol A over a ZnFe2O4/TiO2 nanocomposite under visible light. Sci. Total Environ. 2019, 646, 745–756. [Google Scholar] [CrossRef]
  120. Nguyen, L.T.T.; Nguyen, L.T.H.; Manh, N.C.; Quoc, D.N.; Quang, H.N.; Nguyen, H.T.T.; Nguyen, D.C.; Bach, L.G. A Facile Synthesis, Characterization, and Photocatalytic Activity of Magnesium Ferrite Nanoparticles via the Solution Combustion Method. J. Chem. 2019, 2019, 3428681. [Google Scholar] [CrossRef]
  121. Sun, M.; Han, X.; Chen, S. Synthesis and photocatalytic activity of nano-cobalt ferrite catalyst for the photo-degradation various dyes under simulated sunlight irradiation. Mater. Sci. Semicond. Process. 2019, 91, 367–376. [Google Scholar] [CrossRef]
  122. Sundararajan, M.; John Kennedy, L.; Nithya, P.; Judith Vijaya, J.; Bououdina, M. Visible Light Driven Photocatalytic Degradation of Rhodamine B Using Mg Doped Cobalt Ferrite Spinel Nanoparticles Synthesized by Microwave Combustion Method. J. Phys. Chem. Solids 2017, 108, 61–75. [Google Scholar] [CrossRef]
  123. Singh, G.; Kaur, M.; Garg, V.K.; Oliveira, A.C. Oxygen hyper stoichiometric trimetallic titanium doped magnesium ferrite: Structural and photocatalytic studies. Ceram. Int. 2022, 48, 24476–24484. [Google Scholar] [CrossRef]
  124. Sharma, R.; Komal; Kumar, V.; Bansal, S.; Singhal, S. Boosting the catalytic performance of pristine CoFe2O4 with yttrium (Y3+) inclusion in the spinel structure. Mater. Res. Bull. 2017, 90, 94–103. [Google Scholar] [CrossRef]
  125. Dhiman, M.; Chudasama, B.; Kumar, V.; Tikoo, K.B.; Singhal, S. Augmenting the photocatalytic performance of cobalt ferrite via change in structural and optical properties with the introduction of different rare earth metal ions. Ceram. Int. 2019, 45, 3698–3709. [Google Scholar] [CrossRef]
  126. Sathishkumar, P.; Mangalaraja, R.V.; Anandan, S.; Ashokkumar, M. CoFe2O4/TiO2 nanocatalysts for the photocatalytic degradation of Reactive Red 120 in aqueous solutions in the presence and absence of electron acceptors. Chem. Eng. J. 2013, 220, 302–310. [Google Scholar] [CrossRef]
  127. Barakat, N.A.M.; Kanjwal, M.A.; Chronakis, I.S.; Kim, H.Y. Influence of temperature on the photodegradation process using Ag-doped TiO2 nanostructures: Negative impact with the nanofibers. J. Mol. Catal. A Chem. 2013, 366, 333–340. [Google Scholar] [CrossRef] [Green Version]
  128. Kumar, G.P.A. A Review on the Factors Affecting the Photocatalytic Degradation of Hazardous Materials. Mater. Sci. Eng. Int. J. 2017, 1, 1–10. [Google Scholar] [CrossRef] [Green Version]
  129. Ubhi, M.K.; Kaur, M.; Singh, D.; Sharma, V.K. Structural and adsorptive properties of boron- and phosphorous-doped graphene oxide: Insight into effective removal of Pb (II) and as (III). J. Water Process Eng. 2023, 52, 103539. [Google Scholar] [CrossRef]
  130. Zhao, Y.; Lin, C.; Bi, H.; Liu, Y.; Yan, Q. Magnetically separable CuFe2O4/AgBr composite photocatalysts: Preparation, characterization, photocatalytic activity and photocatalytic mechanism under visible light. Appl. Surf. Sci. 2017, 392, 701–707. [Google Scholar] [CrossRef]
  131. Koe, W.S.; Lee, J.W.; Chong, W.C.; Pang, Y.L.; Sim, L.C. An overview of photocatalytic degradation: Photocatalysts, mechanisms, and development of photocatalytic membrane. Environ. Sci. Pollut. Res. 2020, 27, 2522–2565. [Google Scholar] [CrossRef]
  132. Zafar, M.N.; Dar, Q.; Nawaz, F.; Zafar, M.N.; Iqbal, M.; Nazar, M.F. Effective adsorptive removal of azo dyes over spherical ZnO nanoparticles. J. Mater. Res. Technol. 2019, 8, 713–725. [Google Scholar] [CrossRef]
  133. Gomes, J.; Lincho, J.; Domingues, E.; Quinta-Ferreira, R.M.; Martins, R.C. N–TiO2 Photocatalysts: A Review of Their Characteristics and Capacity for Emerging Contaminants Removal. Water 2019, 11, 373. [Google Scholar] [CrossRef] [Green Version]
  134. Pasinszki, T.; Krebsz, M. Synthesis and Application of Zero-Valent Iron Nanoparticles in Water Treatment, Environmental Remediation, Catalysis, and Their Biological Effects. Nanomaterials 2020, 10, 917. [Google Scholar] [CrossRef]
  135. Miri, A.; Mahabbati, F.; Najafidoust, A.; Miri, M.J.; Sarani, M. Nickel oxide nanoparticles: Biosynthesized, characterization and photocatalytic application in degradation of methylene blue dye. Inorg. Nano-Metal Chem. 2022, 52, 122–131. [Google Scholar] [CrossRef]
  136. Raliya, R.; Avery, C.; Chakrabarti, S.; Biswas, P. Photocatalytic degradation of methyl orange dye by pristine titanium dioxide, zinc oxide, and graphene oxide nanostructures and their composites under visible light irradiation. Appl. Nanosci. 2017, 7, 253–259. [Google Scholar] [CrossRef] [Green Version]
  137. Singh, S.; Kumar, V.; Romero, R.; Sharma, K.; Singh, J. Applications of Nanoparticles in Wastewater Treatment. In Nanobiotechnology in Bioformulations; Prasad, R., Kumar, V., Kumar, M., Choudhary, D., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 395–418. ISBN 9783030170608. [Google Scholar]
  138. Dimapilis, E.A.S.; Hsu, C.-S.; Mendoza, R.M.O.; Lu, M.-C. Zinc oxide nanoparticles for water disinfection. Sustain. Environ. Res. 2018, 28, 47–56. [Google Scholar] [CrossRef]
  139. Dobrucka, R. Facile Synthesis of Trimetallic Nanoparticles Au/CuO/ZnO Using Vitex Agnus-Castus Extract and Their Activity in Degradation of Organic Dyes. Int. J. Environ. Anal. Chem. 2021, 101, 2046–2057. [Google Scholar] [CrossRef]
  140. Kirankumar, V.S.; Sumathi, S. Photocatalytic and antibacterial activity of bismuth and copper co-doped cobalt ferrite nanoparticles. J. Mater. Sci. Mater. Electron. 2018, 29, 8738–8746. [Google Scholar] [CrossRef]
  141. Sun, F.; Zeng, Q.; Tian, W.; Zhu, Y.; Jiang, W. Magnetic MFe2O4-Ag2O (M = Zn, Co, & Ni) composite photocatalysts and their application for dye wastewater treatment. J. Environ. Chem. Eng. 2019, 7, 103011. [Google Scholar] [CrossRef]
  142. Babu, B.; Koutavarapu, R.; Shim, J.; Yoo, K. SnO2 quantum dots decorated NiFe2O4 nanoplates: 0D/2D heterojunction for enhanced visible-light-driven photocatalysis. Mater. Sci. Semicond. Process. 2020, 107, 104834. [Google Scholar] [CrossRef]
  143. Kuang, M.; Zhang, J.; Wang, W.; Chen, J.; Liu, R.; Xie, S.; Wang, J.; Ji, Z. Synthesis of octahedral-like ZnO/ZnFe2O4 heterojunction photocatalysts with superior photocatalytic activity. Solid State Sci. 2019, 96, 105901. [Google Scholar] [CrossRef]
  144. Emadian, S.S.; Ghorbani, M.; Bakeri, G. Magnetically separable CoFe2O4/ZrO2 nanocomposite for the photocatalytic reduction of hexavalent chromium under visible light irradiation. Synth. Met. 2020, 267, 116470. [Google Scholar] [CrossRef]
  145. Silva, E.D.N.; Brasileiro, I.L.O.; Madeira, V.S.; de Farias, B.A.; Ramalho, M.L.A.; Rodríguez-Aguado, E.; Rodríguez-Castellón, E. Reusable CuFe2O4–Fe2O3 catalyst synthesis and application for the heterogeneous photo-Fenton degradation of methylene blue in visible light. J. Environ. Chem. Eng. 2020, 8, 104132. [Google Scholar] [CrossRef]
  146. Munir, S.; Rasheed, A.; Zulfiqar, S.; Aadil, M.; Agboola, P.O.; Shakir, I.; Warsi, M.F. Synthesis, characterization and photocatalytic parameters investigation of a new CuFe2O4/Bi2O3 nanocomposite. Ceram. Int. 2020, 46, 29182–29190. [Google Scholar] [CrossRef]
  147. Lu, D.; Zhang, Y.; Lin, S.; Wang, L.; Wang, C. Synthesis of magnetic ZnFe2O4/graphene composite and its application in photocatalytic degradation of dyes. J. Alloys Compd. 2013, 579, 336–342. [Google Scholar] [CrossRef]
  148. Behera, A.; Mansingh, S.; Das, K.K.; Parida, K. Synergistic ZnFe2O4-carbon allotropes nanocomposite photocatalyst for norfloxacin degradation and Cr (VI) reduction. J. Colloid Interface Sci. 2019, 544, 96–111. [Google Scholar] [CrossRef]
  149. Sudhaik, A.; Raizada, P.; Shandilya, P.; Singh, P. Magnetically recoverable graphitic carbon nitride and NiFe2O4 based magnetic photocatalyst for degradation of oxytetracycline antibiotic in simulated wastewater under solar light. J. Environ. Chem. Eng. 2018, 6, 3874–3883. [Google Scholar] [CrossRef]
  150. Palanivel, B.; Ayappan, C.; Jayaraman, V.; Chidambaram, S.; Maheswaran, R.; Mani, A. Inverse spinel NiFe2O4 deposited g-C3N4 nanosheet for enhanced visible light photocatalytic activity. Mater. Sci. Semicond. Process. 2019, 100, 87–97. [Google Scholar] [CrossRef]
  151. Wang, Q.; Hui, J.; Li, J.; Cai, Y.; Yin, S.; Wang, F.; Su, B. Photodegradation of methyl orange with PANI-modified BiOCl photocatalyst under visible light irradiation. Appl. Surf. Sci. 2013, 283, 577–583. [Google Scholar] [CrossRef]
  152. Deng, Y.; Tang, L.; Zeng, G.; Dong, H.; Yan, M.; Wang, J.; Hu, W.; Wang, J.; Zhou, Y.; Tang, J. Enhanced visible light photocatalytic performance of polyaniline modified mesoporous single crystal TiO2 microsphere. Appl. Surf. Sci. 2016, 387, 882–893. [Google Scholar] [CrossRef]
  153. Zhou, J.; Zhang, Z.; Kong, X.; He, F.; Zhao, R.; Wu, R.; Wei, T.; Wang, L.; Feng, J. A novel P-N heterojunction with staggered energy level based on ZnFe2O4 decorating SnS2 nanosheet for efficient photocatalytic degradation. Appl. Surf. Sci. 2020, 510, 145442. [Google Scholar] [CrossRef]
  154. Huang, S.; Xu, Y.; Zhou, T.; Xie, M.; Ma, Y.; Liu, Q.; Jing, L.; Xu, H.; Li, H. Constructing magnetic catalysts with in-situ solid-liquid interfacial photo-Fenton-like reaction over Ag3PO4@NiFe2O4 composites. Appl. Catal. B Environ. 2018, 225, 40–50. [Google Scholar] [CrossRef]
  155. Li, X.; Wang, L.; Zhang, L.; Zhuo, S. A facile route to the synthesis of magnetically separable BiOBr/NiFe2O4 composites with enhanced photocatalytic performance. Appl. Surf. Sci. 2017, 419, 586–594. [Google Scholar] [CrossRef]
  156. Eskandari, N.; Nabiyouni, G.; Masoumi, S.; Ghanbari, D. Preparation of a new magnetic and photo-catalyst CoFe2O4–SrTiO3 perovskite nanocomposite for photo-degradation of toxic dyes under short time visible irradiation. Compos. Part B Eng. 2019, 176, 107343. [Google Scholar] [CrossRef]
  157. Jia, J.; Du, X.; Zhang, Q.; Liu, E.; Fan, J. Z-scheme MgFe2O4/Bi2MoO6 heterojunction photocatalyst with enhanced visible light photocatalytic activity for malachite green removal. Appl. Surf. Sci. 2019, 492, 527–539. [Google Scholar] [CrossRef]
  158. Wang, J.; Zhang, Q.; Deng, F.; Luo, X.; Dionysiou, D.D. Rapid toxicity elimination of organic pollutants by the photocatalysis of environment-friendly and magnetically recoverable step-scheme SnFe2O4/ZnFe2O4 nano-heterojunctions. Chem. Eng. J. 2020, 379, 122264. [Google Scholar] [CrossRef]
  159. Suresh, R.; Rajendran, S.; Kumar, P.S.; Vo, D.-V.N.; Cornejo-Ponce, L. Recent advancements of spinel ferrite based binary nanocomposite photocatalysts in wastewater treatment. Chemosphere 2021, 274, 129734. [Google Scholar] [CrossRef] [PubMed]
  160. Mir, S.H.; Jennings, B.D.; Akinoglu, G.E.; Selkirk, A.; Gatensby, R.; Mokarian-Tabari, P. Enhanced Dye Degradation through Multi-Particle Confinement in a Porous Silicon Substrate: A Highly Efficient, Low Band Gap Photocatalyst. Adv. Opt. Mater. 2021, 9, 2002238. [Google Scholar] [CrossRef]
  161. Wilson, A.; Mishra, S.R.; Gupta, R.; Ghosh, K. Preparation and photocatalytic properties of hybrid core–shell reusable CoFe2O4–ZnO nanospheres. J. Magn. Magn. Mater. 2012, 324, 2597–2601. [Google Scholar] [CrossRef]
  162. Suwarnkar, M.B.; Dhabbe, R.S.; Kadam, A.N.; Garadkar, K.M. Enhanced photocatalytic activity of Ag doped TiO2 nanoparticles synthesized by a microwave assisted method. Ceram. Int. 2014, 40, 5489–5496. [Google Scholar] [CrossRef]
  163. Zhu, X.; Zhang, F.; Wang, M.; Ding, J.; Sun, S.; Bao, J.; Gao, C. Facile synthesis, structure and visible light photocatalytic activity of recyclable ZnFe2O4/TiO2. Appl. Surf. Sci. 2014, 319, 83–89. [Google Scholar] [CrossRef]
  164. Gupta, V.K.; Eren, T.; Atar, N.; Yola, M.L.; Parlak, C.; Karimi-Maleh, H. CoFe2O4@TiO2 decorated reduced graphene oxide nanocomposite for photocatalytic degradation of chlorpyrifos. J. Mol. Liq. 2015, 208, 122–129. [Google Scholar] [CrossRef]
  165. Masunga, N.; Mmelesi, O.K.; Kefeni, K.K.; Mamba, B.B. Recent advances in copper ferrite nanoparticles and nanocomposites synthesis, magnetic properties and application in water treatment: Review. J. Environ. Chem. Eng. 2019, 7, 103179. [Google Scholar] [CrossRef]
  166. Rani, M.; Rachna; Shanker, U. Efficient photocatalytic degradation of Bisphenol A by metal ferrites nanoparticles under sunlight. Environ. Technol. Innov. 2020, 19, 100792. [Google Scholar] [CrossRef]
Figure 1. Number of publications vs. year. Results obtained from Science Direct database (as of April 2023). Keywords: Photocatalytic activity of ferrite.
Figure 1. Number of publications vs. year. Results obtained from Science Direct database (as of April 2023). Keywords: Photocatalytic activity of ferrite.
Processes 11 01727 g001
Figure 2. Structure of spinel ferrite showing tetrahedral (yellow), octahedral (blue) and oxygen (red) arrangement. Adapted with permission [39]; copyright (2019), Elsevier.
Figure 2. Structure of spinel ferrite showing tetrahedral (yellow), octahedral (blue) and oxygen (red) arrangement. Adapted with permission [39]; copyright (2019), Elsevier.
Processes 11 01727 g002
Figure 4. (a) Effect of Mn0.5Zn0.125Mg0.375Fe2O4 dose on the degradation of Rhodamine B dye. Adapted with permission [100]; copyright (2020), Elsevier. (b) Photocatalytic degradation of methylene blue dye as a function of irradiation time at different CoFe2O4/rGO dosages. Inset shows the value of methylene blue dye degradation efficiency as a function of CoFe2O4/rGO dosages. Adapted with permission [115]; copyright (2020), Elsevier.
Figure 4. (a) Effect of Mn0.5Zn0.125Mg0.375Fe2O4 dose on the degradation of Rhodamine B dye. Adapted with permission [100]; copyright (2020), Elsevier. (b) Photocatalytic degradation of methylene blue dye as a function of irradiation time at different CoFe2O4/rGO dosages. Inset shows the value of methylene blue dye degradation efficiency as a function of CoFe2O4/rGO dosages. Adapted with permission [115]; copyright (2020), Elsevier.
Processes 11 01727 g004
Figure 5. Effect of Zn content on the degradation ofcarbamazepine using ZnFe2O4 NPs. Adapted with permission [118]; copyright (2021), Elsevier.
Figure 5. Effect of Zn content on the degradation ofcarbamazepine using ZnFe2O4 NPs. Adapted with permission [118]; copyright (2021), Elsevier.
Processes 11 01727 g005
Figure 6. Photocatalytic degradation of methylene blue dye as a function of irradiation time at different substrate concentration. Inset shows the value of methylene blue degradation efficiency as a function of substrate concentration. Adapted with permission [115]; copyright (2020), Elsevier.
Figure 6. Photocatalytic degradation of methylene blue dye as a function of irradiation time at different substrate concentration. Inset shows the value of methylene blue degradation efficiency as a function of substrate concentration. Adapted with permission [115]; copyright (2020), Elsevier.
Processes 11 01727 g006
Figure 7. Pathway for photogenerated charges transfer between metal oxide and ferrite under visible light with wide band gap. PC−I: photocatalyst 1 and PC−II: photocatalyst 2.
Figure 7. Pathway for photogenerated charges transfer between metal oxide and ferrite under visible light with wide band gap. PC−I: photocatalyst 1 and PC−II: photocatalyst 2.
Processes 11 01727 g007
Figure 8. (a): Production for free radicals mechanism under visible light by ferrite/reduced graphene oxide nanocomposite. (b): Free radicals generation mechanism under visible light by spinel ferrite/polymer nanocomposite. HOMO−highest occupied molecular orbital and LUMO−lowest unoccupied molecular orbital.
Figure 8. (a): Production for free radicals mechanism under visible light by ferrite/reduced graphene oxide nanocomposite. (b): Free radicals generation mechanism under visible light by spinel ferrite/polymer nanocomposite. HOMO−highest occupied molecular orbital and LUMO−lowest unoccupied molecular orbital.
Processes 11 01727 g008
Figure 9. Free radicals production using S-scheme and Z-scheme mechanisms. Adapted with permission [159]; copyright (2021), Elsevier.
Figure 9. Free radicals production using S-scheme and Z-scheme mechanisms. Adapted with permission [159]; copyright (2021), Elsevier.
Processes 11 01727 g009
Table 1. Advantages and disadvantages of various methods used in the synthesis of ferrites and their nanocomposites.
Table 1. Advantages and disadvantages of various methods used in the synthesis of ferrites and their nanocomposites.
TypeMethodsAdvantagesDisadvantages
PhysicalBall Milling
  • Bulk synthesis
  • Rotation speed can affect the particle size
  • Structural and chemical changes depends on mechanical energy
  • Irregular shape and size
  • Production of contaminants can occur
  • Disturbance in the lattice structure
ChemicalSol-gel
  • Processing temperature is low
  • Porosity and homogeneity
  • Yield of the particles is low
  • Gases are produced in large amount
Co-precipitation
  • Economic
  • Low processing temperature
  • irregularity in particle shape and size
Combustion
  • Processing period is low
  • High yield with purity
  • Production of large amount of gases
Micro emulsion
  • Easy processing
  • Controllable and uniform particle size and homogeneity
  • Particle yield is low
  • liquids are required in large amounts
Microwave and Hydrothermal
  • Yield is high with low processing period
  • Costly equipment
  • Irregularity in particle shape and size
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Singh, G.; Ubhi, M.K.; Jeet, K.; Singla, C.; Kaur, M. A Review on Impacting Parameters for Photocatalytic Degradation of Organic Effluents by Ferrites and Their Nanocomposites. Processes 2023, 11, 1727. https://doi.org/10.3390/pr11061727

AMA Style

Singh G, Ubhi MK, Jeet K, Singla C, Kaur M. A Review on Impacting Parameters for Photocatalytic Degradation of Organic Effluents by Ferrites and Their Nanocomposites. Processes. 2023; 11(6):1727. https://doi.org/10.3390/pr11061727

Chicago/Turabian Style

Singh, Gurpinder, Manpreet Kaur Ubhi, Kiran Jeet, Chetan Singla, and Manpreet Kaur. 2023. "A Review on Impacting Parameters for Photocatalytic Degradation of Organic Effluents by Ferrites and Their Nanocomposites" Processes 11, no. 6: 1727. https://doi.org/10.3390/pr11061727

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop